Next Article in Journal
Machine-Learning-Assisted De Novo Design of Organic Molecules and Polymers: Opportunities and Challenges
Next Article in Special Issue
Tailoring Gellan Gum Spongy-Like Hydrogels’ Microstructure by Controlling Freezing Parameters
Previous Article in Journal
New Monomer Based on Eugenol Methacrylate, Synthesis, Polymerization and Copolymerization with Methyl Methacrylate–Characterization and Thermal Properties
Previous Article in Special Issue
Synergistic Effects on Incorporation of β-Tricalcium Phosphate and Graphene Oxide Nanoparticles to Silk Fibroin/Soy Protein Isolate Scaffolds for Bone Tissue Engineering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Piezoelectric Scaffolds as Smart Materials for Neural Tissue Engineering

by
Angelika Zaszczynska
*,
Paweł Sajkiewicz
and
Arkadiusz Gradys
Institute of Fundamental Technological Research, Polish Academy of Sciences, Pawinskiego 5b St., 02-106 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Polymers 2020, 12(1), 161; https://doi.org/10.3390/polym12010161
Submission received: 29 November 2019 / Revised: 31 December 2019 / Accepted: 5 January 2020 / Published: 8 January 2020
(This article belongs to the Special Issue Biodegradable Polymer Scaffolds for Tissue Engineering)

Abstract

:
Injury to the central or peripheral nervous systems leads to the loss of cognitive and/or sensorimotor capabilities, which still lacks an effective treatment. Tissue engineering in the post-injury brain represents a promising option for cellular replacement and rescue, providing a cell scaffold for either transplanted or resident cells. Tissue engineering relies on scaffolds for supporting cell differentiation and growth with recent emphasis on stimuli responsive scaffolds, sometimes called smart scaffolds. One of the representatives of this material group is piezoelectric scaffolds, being able to generate electrical charges under mechanical stimulation, which creates a real prospect for using such scaffolds in non-invasive therapy of neural tissue. This paper summarizes the recent knowledge on piezoelectric materials used for tissue engineering, especially neural tissue engineering. The most used materials for tissue engineering strategies are reported together with the main achievements, challenges, and future needs for research and actual therapies. This review provides thus a compilation of the most relevant results and strategies and serves as a starting point for novel research pathways in the most relevant and challenging open questions.

1. Introduction

The nervous system is the most complicated system in the body affecting the sensory and motor functions, when the system is damaged. Injuries of the central nervous system (CNS), i.e., brain and spinal cord, lead usually to permanent disability due to severe limitations for spontaneous regeneration of the CNS [1,2,3,4,5,6,7,8,9], leading to considerable socio-economic problems. For instance, 577 cases of traumatic brain injuries (TBI) per 100,000 people per year occurred in the U.S. alone, while, in Europe, the number of patients with diagnosed TBI was estimated at 262 per 100,000 [10]. Spinal cord injuries (SCI) are predominantly associated with irreversible loss of motor functions. In the U.S., 39 per 100,000 people were estimated to be SCI victims every year, mostly due to traffic accidents, jumping in pools, and falling from heights, while in Europe, the amount of SCI cases was 15 per 100,000 per year [11]. Regeneration of damaged neural tissue is often hindered by the presence of internal factors, such as tumor and scar tissue formation, which blocks its reconstruction. Up to date, there is no effective therapy for TBI and SCI, and surgery is able to inhibit lesion spreading only, while drug therapies have been so far focused mostly on pain relief. Hence, the current goal of CNS tissue engineering is to design a biomaterial enabling the effective outgrowth and differentiation of neural stem cells [12,13,14].
In recent decades, there has been increasing interest in research related to the development of smart materials [15]. Such materials are generally designed to respond to external stimuli (physical, chemical, mechanical) and behave similar to natural body tissues. One type of such smart materials is piezoelectric scaffolds, which can generate electrical signals in response to the applied stress [16]. Furthermore, they can stimulate the signaling pathways and thereby enhance the tissue regeneration at the impaired site. This applies especially to neural tissue, where the electrical charges are crucial for cellular activity. The major advantage of such piezoelectric scaffolds is that electrical potential can be generated non-invasively under the influence of mechanical forces, without the need to use invasive electrodes [17,18].
It is known that obtaining piezoelectric scaffolds is possible using various paths, involving solvent casting, TIPS (thermally induced phase separation), freeze drying, or solution blowing [19]. One of most simple and effective methods of fabricating scaffolds in the form of ultrafine fibers with diameters ranging from a few nanometers to several micrometers is electrospinning. This relatively new method uses electric force applied in the form of a very high electrostatic field to draw charged threads from polymer solutions or melts up to the fiber. The process of electrospinning depends on various parameters, which are usually divided into three groups related to the process, material, and ambient parameters. This quite simple and inexpensive technique enables the formation of nano and submicron fibers, the properties of which differ substantially from the ones observed in the bulk materials [20].
For neural tissue engineering, a broad spectrum of synthetic and natural polymers has been studied in the form of electrospun scaffolds [21,22,23,24,25,26,27,28,29,30]. Nerve regeneration is a localized and complex biological phenomenon that makes the treatment of patients suffering from nervous system injuries difficult. Therefore, the application of piezoelectric polymers as nerve guidance conduits allows direct delivery of electrical stimulation of the cell’s ingrowth with its electrical activity during mechanical deformation without the need for an external power source. Recent investigations have proven that neurons are extremely sensitive to electrical signals [30,31,32,33,34,35,36,37,38,39].
The aim of this paper is to summarize the studies on piezoelectric materials used for neural tissue engineering. The most used materials for neural tissue engineering strategies are reported together with the main achievements, challenges, and future needs for research and actual therapies. This review provides thus a compilation of the most relevant results and strategies and serves as a starting point for novel research pathways in the most relevant and challenging open questions.

2. Mechanotransduction and Piezoelectricity in Living Organisms

Mechanotransduction is any of the various mechanisms through which cells sense and convert mechanical stimuli into electrochemical activity. The well known direct element on the way between external stress transmitted through the extracellular matrix and the cells are s.c. stress activated channels, being membrane proteins capable of responding to mechanical stimuli, i.e., opening or closing, generating selective ion fluxes inside the cell, resulting in a cascade of signaling processes [40,41,42,43]. Other types of transmembrane ion channels are voltage gated channels, which are activated by changes in the electrical membrane potential near the channel (Figure 1) [44,45,46]. They play a key role in excitable cells such as neuronal and muscle tissues, allowing rapid and coordinated depolarization in response to triggering voltage change. Found along the axon and at the synapse, voltage gated ion channels directionally propagate electrical signals. Voltage gated ion channels are usually ion specific, for instance to sodium (Na+), potassium (K+), calcium (Ca2+), and chloride (Cl−) ions. In the case of voltage gated channels, the natural piezoelectricity of the body elements is important for the generation of electrical charges under mechanical stress [47,48,49,50].
Mechanotransduction influences many aspects of biological functions. Nerves and neural stem cells (NSC) are sensitive to their surrounding environment, and they interact with this environment through cell surface receptors. Niche features such as substrate bound molecules, extracellular matrix (ECM) proteins, and properties such as stiffness and topography affect cell adhesion, survival, proliferation, migration, morphology, and differentiation [51,52,53,54,55,56,57,58].
The unique features of the nervous system present challenges to bioengineering research addressing nerve injuries. The nervous system is classified into the central nervous system (CNS) containing brain and spinal cord and the peripheral nervous system (PNS) created by the nerves leaving the CNS (Figure 2). The PNS somatic system transmits sensory and motor information for the CNS, while the autonomic one controls automatic functions (e.g., heart beating, blood pressure) [60,61,62].
The basic functional units in the nervous system having specific electrical properties are neurons and neuroglia (Figure 3), which enable effective transmission of the signals [63,64]. The plasma membrane in a non-excited state is characterized by the resting potential, the value of which is usually around −70 mV [65,66,67].
The electrical properties of the neural cells are described by transmission of electrical signals. This phenomenon strongly affects cell behavior via activated ion influx/efflux across the cell membrane. Another phenomenon connected with the transmission of the signals in nerve cells is the action potential. This potential can be dispersed along the axon, which next releases the neurotransmitter (presynaptic ending), and the potential is spread further. A method to activate the action potential is the application of an electrical signal across the neuron. When a dendrite receives an electrical stimulus, the Na+ channels open, and the potential changes from −70 mV to −55 mV. When the potential changes up to +30 mV, the depolarization process starts. Next, the Na+ channels close, and the K+ ones open, completing the depolarization process. When the potential reaches a value around −90 mV, hyperpolarization begins, and the next step is the repolarization of the membrane, which allows receiving another stimulus through the neuron (Figure 4). After the hyperpolarization, the Na+ and K+ channels restore the potential state at the level of −70 mV [68].
Signal transmission is the integral objective of neurons; hence, they are influenced by electrical stimuli. Many research groups have tried to explain the effect of electric stimulation on nerve regeneration. One group described the general effect of electrical stimulation on neurons [70]. Another group studied the activation of the growth controlling transport processes across the plasma membrane and the electrophoretic accumulation of the surface molecules responsible for neurite growth or cell substratum adhesion [70]. Freeman et al. [71] suggested that changes in ionic currents comprise a possible phenomenon that can affect nerve cells, while another thesis [72] described the effect of electrical stimulation on the synthesis of protein and stimulation of the neurite outgrowth in vitro. Furthermore, the authors in [73] postulated that pheochromocytoma in rat neuronal cells (PC12 cells) was electrically activated, while in [74], it was shown that the electrical stimulation increased the adsorption of fibronectin, which explains the enhanced neurite extension on electrically stimulated polypyrrole films. Additionally, nerve cells in the presence of electrical stimulation showed an extensively elongated morphology. In general, tensile/compression forces acting on the piezoelectric scaffolds generate the electrical stimulation and transfer it to the surrounding cells, promoting the cell signaling pathways, responsible for growth factor synthesis (Figure 5) [75].
In 1940, Martin [76] reported the first piezoelectric phenomena in biological tissues, when he observed electric potentials from a bundle of wool compressed by two brass plates. The main component of hair, horns, and wool is keratin in the form of an alpha-helix. It is known that the piezoelectricity of this tissue is due to the highly ordered arrangement and natural polarization of the alpha-helices, which are stabilized by hydrogen bonds between the hydrogen in the amine group and the oxygen in the carbonyl group [77,78,79,80,81,82] (Figure 6).
The results of investigations on the piezoelectric properties of biological tissues have been extensively reported [83,84,85]. The piezoelectric phenomenon has been confirmed in a variety of biological tissues, such as bone and tendons, due to the presence of highly ordered crystalline fibrillar structures such as collagen, chitin, and elastin [86,87,88,89,90].
The piezoelectric effect, by definition, is described by four piezoelectric coefficients d, e.g., and h:
d = (δD/δX)E = (δx/δE)X
e = (δD/δx)E = −(δX/δE)x
g = −(δE/δX)D = (δx/δD)X
h = −(δE/δx)D = −(δX/δD)x
which relate the electrical variables: D (electric displacement) and E (electric field) with the mechanical variables: X (stress) and x (strain). The first terms in Equations (1)–(4) describe the direct piezoelectric effect, while the second terms the converse piezoelectric effect. The subscripts relate to zero constraints of E, D, X, or x. Moreover, each of the coefficients d, e, g, and h is a third rank tensor expressed as a 3 × 6 matrix. The piezoelectric coefficients are indicated in the scheme in Figure 7, along with the coefficients ε (dielectric permittivity) and c (elastic constant) relating to each other the electrical variables D and E as well as the mechanical variables X and x, respectively.
The ratios of the coefficients ε and c at zero constraints of x, X, E, and D, as indicated by superscripts, define the electromechanical coupling coefficient, K, as given by:
εxX = cE/cD = 1 − K2
which characterizes the efficiency of the conversion of mechanical energy to electrical energy and vice versa in the direct and converse piezoelectric effect, respectively. The piezoelectric coefficients may be determined through the direct and converse effects or the piezoelectric resonance, providing directly d, e, g, and h coefficients or the efficiency coefficient K, respectively. The results are complementary, as they cover the low (below audio) and the high frequency ranges (above 10 kHz) [91].
The piezoelectric effect is exhibited in all of the amino acid crystals, excluding alpha glycine alone. The piezoelectric coefficients found in biological materials are generally low, typically in the range of 0.1–10 pm/V (the converse effect) with the values for collagen as low as 0.2–2.0 pC/N (the direct effect). The piezoelectric properties for the most relevant natural materials are reported in Table 1.

3. Scaffolds: Stimuli Responsive (Piezoelectric) vs. Passive

It is generally perceived that the tissue engineering scaffolds should mimic natural existing extracellular matrix (ECM), having similarities as much as possible to the native tissues they are intended to replace in terms of the chemical composition, morphology, physical and mechanical properties, as well as biocompatibility and biodegradability. In detail, the fundamental requirements, that need to be taken into account during scaffold designing are: scaffold biocompatibility, appropriate degradation time in the case of biodegradable materials, the presence of interconnected pores in an appropriate size range, scaffold thickness, mechanical properties, and convenience of use during a surgical procedure [95,96].
It is well known that ECM is a highly dynamic structure; it is constantly being remodeled, either enzymatically or non-enzymatically, and its molecular components are subjected to various modifications. However, most of the recent artificial scaffolds remain passive, i.e., non-stimuli responsive to the external changes of the environment (static scaffolds) [97]. Moreover, conventional static scaffolds, even conductive, largely disturb the natural signaling pathways, due to their rigidity towards the signal conduction. Thus, there is a high need for smart, stimuli responsive scaffolds, which can generate and transfer the bioelectric signals analogously to the native tissues for appropriate physiological functions. Piezoelectric materials can generate electrical signals in response to the applied stress, which can be imposed even by attachment and migration of cells or body movements [98]. Using piezoelectric materials as tissue engineering scaffolds enables electrical stimulation without the need for electrodes, an external source of electricity, or implanting batteries. Such scaffolds should possess proper architecture and mechanical properties in addition in order to support cell adhesion, proliferation, and differentiation. The size of pores should be controllable and adjusted in order to enable diffusion of the metabolite, but also for appropriate cell adhesion to the biomaterial. Ninety percent porosity and a pore size in the range of 10–100 µm seem to be the most suitable for neuron growth. Scaffolds characterized by 85–90% porosity may be obtained by the electrospinning technique [99]. Such porosity is reported as supporting cellular migration and controlled diffusion of cells, metabolites, and medium, being important for cells organization, differentiation, and survival [100,101].
Obtaining a scaffold that would be biocompatible, biodegradable, conducting, and resistant to infection in order to provide neurite outgrowth is a complex task [102]. Successful nerve regeneration requires tissue engineered scaffolds not only for mechanical support of growing neurites and impediment of the ingrowth of fibrous scar tissues, but also to send biological signals to guide the axonal growth cone to the distal stump. Polymers are in general that materials that have been extensively used for creating suitable scaffolds for neural tissue [103].
Table 2 summarizes the main works on the application of piezoelectric potential scaffold materials and their applications.

4. Application of Piezoelectric Biomaterials in Neural Tissue Engineering

4.1. Piezoceramics

The earliest studied piezoelectric material group is the piezoceramics. The first applications were dated since around 1950, and since then, they have been widely used in the industry [133]. Wersing et al. [134] conducted the pioneering study on porous piezoceramics, as well as provided the basics in the theory and initial measurements [135]. Currently, there is a great need for lead-free piezoelectric materials, but the most practical ceramics are still based on lead zirconate titanate. Rat cortical neurons cultured on PZT slides coated with poly-L-lysine grew significantly longer axons, despite a decrease in cell number. Furthermore, the frequency and amplitude of the excitatory postsynaptic currents increased, suggesting that piezoelectricity could have augmented neuronal activity [136]. It is worth mentioning that piezoceramics are used for medical applications, especially medical actuators, transducers, and sensors. Due to allergic reactions, piezoceramics are not used in pure solution for medical implants. The innovative systems for medical applications are composites based on polymer matrices, with ceramic fillers, in the form of fibers [137,138,139]. To achieve bone defect repair, Lopes et al. [140] incorporated barium titanate nanoparticles into polymer matrix, which induced relatively high spontaneous polarization. Moreover, this system is characterized by reduced fragility and can be used as electroactive scaffolds [141].

4.1.1. Barium Titanate

The first piezoelectric effects in ceramics were discovered during poling of BT and led to the wide use of this material group, also as an addition in scaffolds, especially in medical applications [142,143]. Piezoceramics based on barium titanate exhibit low toxicity compared to lead based piezoelectric materials. For their high strain, they are among the most investigated groups of piezoceramics. BT nanoparticles have demonstrated cytocompatibility, even at higher concentrations of 100 μg/mL [144]. Ciofani et al. [145] demonstrated that PLGA matrix with the addition of BT nanoparticles supports the cell proliferation and attachment of osteocytes and osteoblasts. Additionally, the incorporation of barium titanate nanoparticles into the polymeric matrix improves the mechanical properties of the composite scaffold [146] and promotes cellular activity in tissue engineering applications [147].

4.1.2. Boron Nitride

Boron nitride (BN) based nanomaterials play a significant role in nanotechnology owing to their conductivity, mechanical strength, and high thermal stability [148,149]. The most known boron nitride piezo-materials are in the shape of nanotubes, and with increased cytocompatibility, they can be used in tissue engineering [150] and drug delivery, due to their high piezoelectric properties [151,152,153,154,155]. It has been proven that boron nitride nanotubes have a positive influence on the adhesion of cells [156]. Among all the properties of boron nitride nanotubes (BNNTs), their excellent piezoelectricity is the most important for using them as nanovectors to deliver electrical or mechanical signals within cells [157].

4.1.3. Zinc Oxide

Zinc oxide based piezoceramics are widely used due to their asymmetric hexagonal wurtzite structure and polar crystal surface. They have found application as piezoelectric nanogenerators, because of the easy fabrication [158]. ZnO in the shape of nanostructures is biocompatible [159]. It has been suggested that with ZnO size, its cytotoxicity increases, which has an influence on the levels of reactive oxygen species, reduces the mitochondrial membrane potential, and induces the production of interleukin in human cells. Additionally, it has been reported that chemical modification can reduce toxicity, providing a way for use in biomedical applications [160,161].

4.2. Piezopolymers

Piezoelectric polymers are a relatively new class of materials allowing the formation of electrical charges under mechanical stimulation in the absence of additional energy sources or electrodes [162]. Additionally, which is very important from the biomedical point of view, polymers are able to meet the requirements of biocompatibility and biodegradability, which is very crucial for new types of implants in regenerative medicine [163]. Further, their big advantage is the very high processing flexibility, which differentiates them from inorganic materials [164].

4.2.1. Synthetic Polymers

Polyvinylidene Fluoride

Among various piezoelectric polymers, PVDF is widely investigated, primarily because of its high piezoelectricity, processability, good chemical resistance, thermal stability, and good mechanical properties as compared to other piezoelectric polymers. PVDF may exist in at least five crystalline polymorphic phases, among which the β-phase shows the highest piezoelectricity, which reaches 20 pC/N (Figure 8) [165,166].
PVDF macromolecules may take various chain conformations and arrangement of CH2–CF2 molecular dipoles, resulting in various net dipole moments. A strong electric moment in the PVDF monomer unit arises from the strong electro-negativity of fluorine atoms as compared to hydrogen atoms. In the case that polymer chains are packed into crystals to form parallel dipoles, the crystal has a non-zero net dipole moment [89,90,91,92,93,94]. Such a molecular arrangement is observed in the β, γ, and δ phases, the first one showing the strongest dipole moment, due to the all-trans conformation. In the case of other chain conformations: TGTG- and T3GT3G-, parallel dipole moment arrangement, as in the δ and γ phases, respectively, leads to lower polarity; in the case of the same conformations, antiparallel chain dipole arrangement leads to the zero net dipole moment as in the α and ϵ phases [167].
The piezoelectricity of PVDF is phase content dependent, which hinges on the processing conditions. Obtaining a particular crystal phase is possible using various paths, involving melt or solution crystallization, annealing (also at high pressure), mechanical drawing, or electrical poling. The presence of polar phases is very important, in particular, due to its bioelectrical effect of the stimulation of the nervous system, holding promise for effective tissue regeneration [168,169].
As regards the most polar β-phase, it may be obtained, for example, by annealing at very high pressure from the α-phase, by poling at a very high electrical field from the α-phase or δ-phase [170] or drawing from the γ-phase [171]. In order to increase the content of polar phases, various methods are reported: melt-recrystallization [172], poling under a high electric field [173], application of high pressure [174], mechanical stretching [175], and the addition of nanoparticles, graphene, and nanowires [176].
The desire to use PVDF in piezoelectric scaffolds in tissue engineering requires the use of fabrication techniques that allow obtaining the proper morphology and high polar phase content responsible for the high piezoelectricity [177]. One of the promising fabrication techniques to fulfil both expectations is the electrospinning technique. Many publications were devoted to electrospinning of PVDF nanofibers from solution [178], determining the effect of processing parameters on the structure and properties of nanofibers and the characteristics of nonwoven nanofiber [179]. The content of the β-phase in PVDF was studied from the point of view of the applied voltage and rotation speed of the rotational collector. The collector rotational speed relates to the mechanical deformation, which is known to promote the formation of the polar phase [180]. Liu et al. [181] formed nanofibers with different rotational speeds of the collector: 900, 1100, 1300, 1500, 1700 and 1900 rpm. XRD diffraction showed a peak around 20.6–20.9 deg, which belonged to the β-phase, while the α-phase peaks disappeared. They received piezoelectric PVDF fibers with a small diameter, smooth surface morphology, and appropriate β-phase at a velocity of 1900 rpm. Recent works support the view that increasing of the rotational speed of the collector induces a higher content of the piezoelectric β-phase [182,183,184].
The hydrophobicity of PVDF is a problematic issue in neural tissue engineering. In order to reduce it, numerous research works have been conducted. In order to enhance the hydrophilic, as well as mechanical and electrical properties, PVDF has been modified by the addition of different nanostructures: nanoparticles [185,186,187], inorganic nanoparticles [188,189], nanotubes [190], and also by the addition of different polymers such as polyethylene glycol (PEG) [191] and polyvinyl alcohol (PVA) [192]. The addition of nanoparticles, especially metallic ones, improves the chemical, physical, and optical properties [193], while diamond nanoparticles have no significant impact on neuroblastoma cell morphology [194]. The incorporation of these nanostructures into a polymer piezoelectric scaffold can positively affect the nerve tissue. Additionally, modification of the surface can increase the neuron length and number of synaptic connections [195,196].
Arinzeh et al. [113,114] tested the piezoelectric scaffolds’ potential for promoting in vitro neural differentiation of human neural stem cells, thus demonstrating their applicability in neural tissue engineering. The authors in [197] extended the above studies by applying mechanical vibration, while generating electric fields to induce the piezoelectric effect in piezopolymers. The activation of the piezoelectric effect can be achieved by choosing various sources of mechanical stimulation, including vibration plates, sound, and ultrasound (US) [198,199]. Hoop et al. [85] investigated the influence of the piezoelectric PVDF substrate on supporting neural differentiation under dynamic stimulation. The results showed that the applied ultrasonic waves were sufficient to induce polarization in piezoelectric PVDF sheets and resulted in differentiation of PC12 cells. Piezoelectric PVDF can influence neuronal differentiation and neurite outgrowth of mouse neuroblastoma cells [200]. Electric fields have been shown to influence the growth and orientation of neurons in vitro, whereas the electric field was generated via electrodes [201]. Other studies have been reported successful neural stimulation in various piezoelectric systems, especially piezoelectric micro- and nano-fibers [202,203].
It was shown that the long term application of piezoelectric stimulation on neurons induces the number, length, and branching of neural cells with respect to non-stimulation conditions. No effects on neurite regeneration were observed when vibrations were applied to non-piezoelectric materials (e.g., mechanical stimulation of neurons) [204,205,206].

Poly-Vinylidene Fluoride-Trifluoroethylene

Among the piezoelectric materials, this copolymer demonstrated the highest electroactive properties with a piezoelectric coefficient as high as 30 pC/N [207]. PVDF-TrFE forms the β-phase through copolymerization without the need for mechanical stretching or drawing [208]. In the case of additional annealing, mechanical stretching, or electrical poling, it is possible to further increase the crystallinity and alignment of the CF2 dipoles, thereby inducing higher piezoelectricity as compared to homopolymer PVDF, dependent on the TrFE content [209]. Electrospun PVDF-TrFE fibrous scaffolds showed higher crystallinity and β-phase content as compared to the starting powder material for neural and bone tissue engineering [101,102,103,104]. PVDF-TrFE and barium titanate piezoelectric composite membrane has been reported as a charge generator to promote bone regeneration [106].
PVDF-TrFE piezoelectric fibrous scaffolds were used to study their influence on neural repair. Many investigations reported a positive influence of PVDF-TrFE scaffolds on nerve cell growth and differentiation [210]. Lee et al. [114] fabricated a PVDF-TrFE piezoelectric electrospun scaffold with different orientations of the fibers, randomly and aligned. It was shown that the scaffold with aligned fibers had the highest potential in neural tissue engineering, especially in neurite outgrowth of dorsal root ganglion neurons. It was observed that PVDF-TrFE scaffolds can promote the formation of mature neural cells exhibiting neuron-like characteristics, while aligned fibers can promote primary neuron extension and can direct the neurite outgrowth [211].
Nerve guidance channels may be built using PVDF-TrFE for neural regeneration [212]. In this study, poled (negatively charged and positively charged) and unpoled channels were used. After four weeks, it was observed that the positively poled channels increased the number of regenerated nerves.
In muscle regeneration, the charge at the surface of PVDF films influences the cell proliferation [213]. However, until now, studies with specific dynamic conditions for piezoelectric PVDF-TrFE with mechanical or electrical stimulation have not been conducted.

Poly-3-Hydroxybutyrate-3-Hydroxyvalerate

PHBV is a polyester with a low piezoelectric coefficient (1.2 pC/N) [214]. It is a thermoplastic produced by many bacteria as an intracellular reservoir of carbon and energy. This polyester is also biodegradable, biocompatible, and exhibits strong mechanical properties, which allow using PHBV as a scaffold in biomedicine and as a biosensor [215,216]. PHBV has a comparable piezoelectric potential to that of bone, which can facilitate bone growth and healing [217,218]; thus, it can be used in the form of a composite with the addition of hydroxyapatite for bone tissue engineering. PHBV has been studied for neural tissue engineering, as a support for neuronal cell growth and axon dendrite polarization [219]. In the form of electrospun aligned PHBV fibers, with the addition of collagen, it can be used as a substrate for nerve tissue engineering [220,221,222].

Poly-L-Lactic Acid

Poly-L-lactic acid is a biodegradable and biocompatible polymer, with a piezoelectric coefficient of −10 pC/N [223]. Fukada et al. demonstrated that implantation of PLLA can promote bone growth in the response of its piezoelectric polarization [93]. PLLA with a structure similar to natural ECM may be used as a biomaterial in various biomedical applications [224]. Aligned PLLA nanofibrous scaffolds coated with graphene oxide promote neural cell growth [225]. Finally, the addition of iron oxide nanoparticles supports extending neurites along electrospun PLLA microfibers [226,227].

4.2.2. Natural Biopolymers

Natural polymers are gaining more importance in tissue engineering because of their biodegradability and low toxicity. In general, many biopolymers exhibit piezoelectricity. As an example, we show some polysaccharides and proteins with relatively strong piezoelectricity.

Cellulose

Cellulose with a piezoelectric coefficient of 0.10 pC/N is a widely investigated natural polymeric material. Cellulose is a linear homopolymer of glucose with high biocompatibility [228,229,230,231,232,233,234]. It is used in different shapes and forms: membrane sponges, microspheres, and non-woven, woven, or knitted textiles. Cellulose has been investigated in the tissue engineering of bones [235,236], cartilage [237], for connective tissue formation [238], as a drug delivery system [239], and as a scaffold for growing functional cardiac cell constructs in vitro [240]. One of the important derivatives of cellulose is methylcellulose (MC), presenting in general good solubility in water, particularly at low temperatures, being dependent on the degree of methyl substitution and the distribution of methoxy groups. Collectively, these data indicate that MC is well suited as a biocompatible injectable scaffold for the repair of brain defects [241,242,243]. Gelatin coated nanoparticles contained in cellulose acetate/PLA scaffolds showed higher cell viability than uncoated scaffolds, and they acted as a nerve guidance conduit for sciatic nerve defects in vitro and in vivo [244], while a gelatin/chitosan/PEDOT hybrid scaffold enhanced the neurite growth of PC12 cells and promoted neuron-like cell adhesion and proliferation [245].

Chitin and Chitosan

Chitin is a natural polysaccharide with a piezoelectric structure with a low piezoelectric coefficient in the range from 0.2 to 1.5 pC/N [246]. It is a natural component of the cuticles of crustaceans, insects, and mollusks. Since it is hydrophilic and biocompatible, chitin is used for biomedical applications, promoting cell adhesion, proliferation, and differentiation [247].
Chitosan is a biodegradable and biocompatible linear polysaccharide obtained by partial deacetylation of chitin. It has been extensively investigated for the preparation of porous scaffolds for cartilage tissue engineering [248]. However, the low mechanical properties of scaffolds prepared from chitosan make its clinical application problematic. An effective method to overcome chitosan’s drawbacks is to blend it with synthetic polymers [249,250,251,252]. Skop et al. designed biocompatible chitosan microspheres for the delivery of neural stem cells and growth factors for CNS injuries [253]; another group designed chitosan particles loaded with the drug piperine, reported to have neuroprotective potential against Alzheimer’s disease, which successfully targeted specific areas of the brain [254]. Chitosan nanoparticles have also been developed for intranasal delivery of therapeutic agents to the brain [255,256]. Aligned PCL/chitosan fibers supported PC12 cells adhesion and growth, enhancing neurite extension along the fiber orientation [257]. PLGA/chitosan scaffolds guided neuronal differentiation for peripheral nerve regeneration both in vitro and in vivo [258,259].

Collagen

This is a natural piezoelectric material with a piezoelectric coefficient in the range from 0.2 to 2.0 pC/N [260]. Research has been reported on the application of collagen scaffolds in bone healing [261,262,263,264]. Furthermore, collagen-calcium phosphate composites have been reported for cartilage tissue engineering [265]. Similarly, collagen-hydroxyapatite piezoelectric composite scaffolds have been proven to be suitable for cellular growth [266]. Collagen scaffolds with the addition of chitosan have been tested in adipose tissue regeneration. Adipocytes were seeded, and the in vitro cytocompatibility and in vivo biocompatibility of scaffolds was confirmed experimentally [267]. An interesting application of collagen is entubulation, hence the use of magnetically aligned type I collagen gel, achieved by exposing the forming collagen gel to a high strength magnetic field, as a filler for collagen tubes. This method was successful in small peripheral nerve lesions, improving nerve regeneration significantly in a 6 mm nerve gap in mice [268] and guiding neurite elongation and Schwann cell invasion in vitro [269] and in vivo [270].

5. Conclusions and Future Perspectives

Recently, smart materials have been of great interest for scientists and physicians, because of the many opportunities to use them as candidates for developing the next generation of biomedical devices, transient implants, and drug delivery vehicles. Considering smart scaffolds for tissue repair and regeneration, piezoelectric materials have recently been of particular interest as they can deliver electrical stimulus without an external power source. There is no doubt that the bioelectric signals produced by piezoelectric scaffolds can regenerate and repair the tissues by definite pathways similar to the natural processes occurring within the natural extracellular matrix (ECM). The combination of morphology together with the chemical, mechanical, and electrical properties of the scaffolds is crucial for the success in tissue regeneration. Electrical charges are particularly important in neural tissue engineering, in which electric pulses can stimulate neurite directional outgrowth to fill gaps in nervous tissue injuries. There is no doubt that the perspective of the broader application of piezoelectric scaffolds as smart materials for neural tissue regeneration is of great importance, allowing avoiding traditional (invasive) electrical stimulation. It was shown recently in in vitro conditions that the deformation of the piezoelectric scaffolds either by mechanical or ultrasound stimulation led to neurite extension and enhanced cell adhesion and proliferation. However, one should be aware that most of the up-to-date experiments using piezoelectric scaffolds were performed without such stimulation, which does not lead to piezoelectricity and resulting electrical charges. In such a case, the only charges that can be active from the cellular perspective are surface charges due to permanent polarization, as well as related to transient deformation caused by the contraction and protrusion of the attached cells. Nonetheless, it is crucial from the perspective of experiments with piezoelectric scaffolds to mimic the in vivo conditions with internal macro- and micro-deformations by in vitro conditions using mechanical (ultrasounds) agitation, allowing obtaining a real piezoelectric response. The next problem in the area of piezoelectric scaffolds is related to the non-biodegradability of the polymers exhibiting the highest piezoelectric coefficients, i.e., PVDF and its copolymers. Therefore, attention should be focused on biodegradable piezoelectric polymers like PHB or PLLA. An interesting alternative, which should be explored in the future, is related to the composite scaffolds containing an electro-conductive polymer like PANi in addition to a piezoelectric polymer. It was shown that the addition of an electro-conductive polymer to the piezoelectric matrix resulted in an increase in piezoelectricity. This kind of composite scaffold should be taken into account, when thinking about biodegradable piezoelectric polymers with originally relatively low piezoelectricity.

Author Contributions

Conceptualization, A.Z., P.S., and A.G.; validation, P.S. and A.G.; writing, original draft preparation, A.Z.; writing, review and editing, A.Z., P.S., and A.G.; visualization, A.Z., P.S., and A.G.; supervision, P.S., and A.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wolfenson, H.; Yang, B.; Sheetz, M.P. Steps in mechanotransduction pathways that control cell morphology. Annu. Rev. Physiol. 2019, 81, 585–605. [Google Scholar] [CrossRef] [PubMed]
  2. Whalin, M.K.; Arora, S.S. Anatomy of the Brain and Spinal Cord. In Basic Sciences in Anesthesia; Farag, E., Argalious, M., Tetzlaff, J., Sharma, D., Eds.; Springer: New York, NY, USA, 2018; pp. 41–59. [Google Scholar]
  3. Lis, A.; Szarek, D.; Laska, J. The outlook for the use of polymeric scaffolds in the reconstruction and the regeneration stimulation of traumatic brain injuries. Polim. Med. 2018, 43, 302–312. [Google Scholar]
  4. Han, D.; Cheung, K.C. Biodegradable cell-seeded nanofiber scaffolds for neural repair. Polymers 2011, 3, 1684–1733. [Google Scholar] [CrossRef]
  5. Balint, R.; Cassidy, N.J.; Cartmell, S.H. Conductive polymers: Towards a smart biomaterial for tissue engineering. Acta Biomat. 2014, 10, 2341–2353. [Google Scholar] [CrossRef]
  6. Niemczyk, B.; Sajkiewicz, P.; Gradys, A. Injectable hydrogels as novel materials for central nervous system regeneration. J. Neural Eng. 2018, 15, 051002. [Google Scholar] [CrossRef]
  7. Lis, A.; Szarek, D.; Laska, J. Biomaterials engineering strategies for spinal cord regeneration: State of the art. Polim. Med. 2013, 43, 59–80. [Google Scholar]
  8. Taylor, C.A.; Bell, J.M.; Breiding, M.J.; Xu, L. Traumatic brain injury—Related emergency Department Visits, Hospitalizations, and Deaths—United States, 2007 and 2013. MMWR Surveill. Summ. 2017, 66, 1. [Google Scholar] [CrossRef]
  9. Mathieu, S.; Manneville, J.B. Intracellular mechanics: Connecting rheology and mechanotransduction. COCEBI 2019, 56, 34–44. [Google Scholar] [CrossRef]
  10. Delcroix, G.J.R.; Schiller, P.C.; Benoit, J.P.; MonteroMenei, C.N. Adult cell therapy for brain neuronal damages and the role of tissue engineering. Biomaterials 2010, 31, 2105–2120. [Google Scholar] [CrossRef]
  11. Wang, S.; Hou, J.; Bei, J.; Zhao, Y. Tissue engineering and peripheral nerve regeneration (III)—Sciatic nerve regeneration with PDLLA nerve guide. Sci. China 2001, 44, 419–426. [Google Scholar] [CrossRef]
  12. Adrian, H.; Mårten, K.; Salla, N.; Lasse, V. Biomarkers of Traumatic Brain Injury: Temporal Changes in Body Fluids. eNeuro 2016, 3, 0294-16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Weidner, N.; Rudiger, R.; Tansey, K.E. Neurological Aspects of Spinal Cord Injury; Springer: Cham, Switzerland, 2017; Volume 1, pp. 3–17. [Google Scholar]
  14. Zhang, L.; Sirivisoot, S.; Balasundaram, G.; Webster, T.J. Advanced Biomaterials: Fundamentals, Processing and Applications; Basu, B., Katti, D., Kumar, A., Eds.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2009. [Google Scholar]
  15. Langlois, J.A.; Rutland-Brown, W.; Thomas, K.E. Traumatic Brain Injury in the United States: Emergency Department Visits, Hospitalizations, and Deaths, Centers for Disease Control and Prevention; National Center for Injury Prevention and Control: Atlanta, GA, USA, 2004.
  16. Flax, J.D.; Aurora, S.; Yang, C. Engraftable human neural stem cells respond to developmental cues, replace neurons, and express foreign genes. Nature Biotechnol. 1998, 16, 1033–1438. [Google Scholar] [CrossRef] [PubMed]
  17. Kordower, J.H.; Tuszynski, M.H. CNS Regeneration: Basic Science and Clinical Advances; Kordower, J.H., Tuszynski, M.H., Eds.; Academic: Cambridge, MA, USA, 1999; pp. 159–182. [Google Scholar]
  18. Steel, E.M.; Sundararaghavan, H.G. Electrically Conductive Materials for Nerve Regeneration. In Neural Engineering; Springer: New York, NY, USA, 2016; pp. 145–179. [Google Scholar]
  19. Zhao, P.; Gu, H.; Mi, H.; Rao, C.; Fu, J.; Turng, L.S. Fabrication of scaffolds in tissue engineering: A review. Front. Mech. Eng. 2018, 13, 107–119. [Google Scholar] [CrossRef]
  20. Gao, S.; Tang, G.; Hua, D.; Xiong, R.; Han, J.; Jiang, S.; Huang, C. Stimuli-responsive bio-based polymeric systems and their applications. J. Mater. Chem. B 2019, 7, 709–729. [Google Scholar] [CrossRef]
  21. Jalili-Firoozinezhad, S.; Mirakhori, F.; Baharvand, H. Nanotissue Engineering of Neural Cells. Stem Cell Nanoeng. 2015, 265, 265–283. [Google Scholar]
  22. Nguyen, H.T.; Wei, C.; Chow, J.K.; Nguy, L.; Nguyen, H.K.; Schmidt, C.E. Electric field stimulation through a substrate influences Schwann cell and extracellular matrix structure. J. Neural Eng. 2013, 10, 046011. [Google Scholar] [CrossRef] [PubMed]
  23. Chew, S.Y.; Wen, Y.; Dzenis, Y.; Leong, K.W. The role of electrospinning in the emerging field of nanomedicine. Curr. Pharm. Des. 2006, 12, 4751–4770. [Google Scholar] [CrossRef] [Green Version]
  24. Murugan, R.; Ramakrishna, S. Nano-featured scaffolds for tissue engineering: A review of spinning methodologies. Tissue Eng. 2006, 12, 435–447. [Google Scholar] [CrossRef]
  25. Venugopal, J.; Low, S.; Choon, A.T.; Ramakrishna, S. Interaction of cells and nanofiber scaffolds in tissue engineering. J. Biomed. Mater. Res. B 2008, 84, 34–48. [Google Scholar] [CrossRef]
  26. Teo, W.E.; He, W.; Ramakrishna, S. Electrospun scaffold tailored for tissue-specific extracellular matrix. Biotechno. J. 2006, 1, 918–929. [Google Scholar] [CrossRef]
  27. Barnes, C.P.; Sell, S.A.; Boland, E.D.; Simpson, D.G.; Bowlin, G.L. Nanofiber technology: Designing the next generation of tissue engineering scaffolds. Adv. Drug Deliv. Rev. 2007, 59, 1413–1433. [Google Scholar] [CrossRef]
  28. Murugan, R.; Ramakrishna, S. Design strategies of tissue engineering scaffolds with controlled fiber orientation. Tissue Eng. 2007, 13, 1845–1866. [Google Scholar] [CrossRef] [PubMed]
  29. Heydarkhan-Hagvall, S.; Schenke-Layland, K.; Dhanasopon, A.P.; Rofail, F.; Smith, H.; Wu, B.M.; Shemin, R.; Beygui, R.E.; MacLellan, W.R. Three-dimensional electrospun ECM-based hybrid scaffolds for cardiovascular tissue engineering. Biomaterials 2008, 29, 2907–2914. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Li, W.J.; Mauck, J.A.; Cooper, X.; Yuan, R.S. Tuan Engineering controllable anisotropy in electrospun biodegradable nanofibrous scaffolds for musculoskeletal tissue engineering. J. Biomech. 2007, 40, 1686–1693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Li, W.J.; Tuli, R.; Huang, X.; Laquerriere, P.; Tuan, R.S. Multilineage differentiation of human mesenchymal stem cells in a three-dimensional nanofibrous scaffold. Biomaterials 2005, 26, 5158–5166. [Google Scholar] [CrossRef] [PubMed]
  32. Yang, F.; Xu, C.Y.; Kotaki, M.; Wang, S.; Ramakrishna, S. Characterization of neural stem cells on electrospun poly(L-lactic acid) nanofibrous scaffold. J. Biomater. Sci. Polym. Ed. 2004, 15, 1483–1497. [Google Scholar] [CrossRef] [PubMed]
  33. Ting, Y.; Gunawan, H.; Sugondo, A.; Chiu, C. A New Approach of Polyvinylidene Fluoride (PVDF) Poling Method for Higher Electric Response. Ferroelectrics 2013, 446, 28–38. [Google Scholar] [CrossRef]
  34. Dang, Z.M.; Lin, Y.H.; Nan, C.W. Novel ferroelectric polymer composites with high dielectric constants. Adv. Mater. 2003, 15, 1625–1629. [Google Scholar] [CrossRef]
  35. Bera, B.; Sarkar, M. Piezoelectricity in PVDF and PVDF Based Piezoelectric Nanogenerator: A Concept. Int. J. Appl. Phys. 2017, 9, 95–99. [Google Scholar] [CrossRef]
  36. Damaraju, S.M.; Wu, S.; Jaffe, M.; Arinzeh, T.L. Structural changes in PVDF fibers due to electrospinning and its effect on biological function. Biomed. Mater. 2013, 8, 045007. [Google Scholar] [CrossRef]
  37. Defteralı, Ç.; Verdejo, R.; Majeed, S.; Boschetti-de-Fierro, A.; Méndez-Gómez, H.R.; Díaz-Guerra, E.; Vuluga, D. In vitro evaluation of biocompatibility of uncoated thermally reduced graphene and carbon nanotube-loaded PVDF membranes with adult neural stem cell-derived neurons and glia. Front. Bioeng. Biotechnol. 2016, 4, 94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Young, T.H.; Lin, U.H.; Lin, D.J.; Chang, H.H.; Cheng, L.P. Immobilization of L-lysine on microporous PVDF membranes for neuron culture. J. Biomater. Sci. Polym. 2009, 20, 703–720. [Google Scholar] [CrossRef] [PubMed]
  39. Bar, H.N.; Bhat, M.R.; Murthy, C.R.L. Identification of failure modes in GFRP using PVDF sensors: ANN approach. Compos. Struct. 2004, 65, 231–237. [Google Scholar] [CrossRef]
  40. Fu, Y.S.; Shih, Y.T.; Cheng, Y.C.; Min, M.Y. Transformation of human umbilical mesenchymal cells into neurons in vitro. J. Biomed. Sci. 2004, 11, 652–660. [Google Scholar] [CrossRef]
  41. Ning, C.; Zhou, Z.; Tan, G.; Zhu, Y.; Mao, C. Electroactive polymers for tissue regeneration: Developments and perspectives. Prog. Polym. Sci. 2018, 81, 144–162. [Google Scholar] [CrossRef]
  42. Moran, H.; Cancel, L.M.; Mayer, M.A.; Qazi, H.; Munn, L.L.; Tarbell, J.M. The cancer cell glycocalyx proteoglycan glypican-1 mediates interstitial flow mechanotransduction to enhance cell migration and metastasis. Biorheology 2019, 56, 151–161. [Google Scholar] [CrossRef]
  43. Gargalionis, A.N.; Basdra, E.K.; Papavassiliou, A.G. Polycystins and Mechanotransduction in Human Disease. Int. J. Mol. Sci. 2019, 20, 2182. [Google Scholar] [CrossRef] [Green Version]
  44. Maurer, M.; Lammerding, J. The driving force: Nuclear mechanotransduction in cellular function, fate, and disease. Annu. Rev. Biomed. Eng. 2019, 21, 443–468. [Google Scholar] [CrossRef]
  45. Yamada, K.M.; Sixt, M. Mechanisms of 3D cell migration. Nat. Rev. Mol. 2019, 20, 738–752. [Google Scholar] [CrossRef]
  46. Chang, W.; Gu, J.G. Impairment of tactile responses and Piezo channel mechanotransduction in mice following chronic vincristine treatment. Neurosci. Lett. 2019, 705, 14–19. [Google Scholar] [CrossRef]
  47. Salvi, A.M.; DeMali, K.A. Mechanisms linking mechanotransduction and cell metabolism. COCEBI 2018, 54, 114–120. [Google Scholar] [CrossRef] [PubMed]
  48. Pariy, I.O.; Ivanova, A.A.; Shvartsman, V.V.; Lupascu, D.C.; Sukhorukov, G.B.; Ludwig, T.; Surmenev, R.A. Piezoelectric Response in Hybrid Micropillar Arrays of Poly (Vinylidene Fluoride) and Reduced Graphene Oxide. Polymers 2019, 11, 1065. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Ruan, L.; Yao, X.; Chang, Y.; Zhou, L.; Qin, G.; Zhang, X. Properties and Applications of the β Phase Poly (vinylidene fluoride). Polymers 2018, 10, 228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Guilak, F.; Cohen, D.M.; Estes, B.T.; Gimble, J.M.; Liedtke, W.; Chen, C.S. Control of stem cell fate by physical interactions with the extracellular matrix. Cell Stem Cell 2009, 5, 17–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Lee, M.R.; Kwon, K.W.; Jung, H.; Kim, H.N.; Suh, K.Y.; Kim, K.; Kim, K.S. Direct differentiation of human embryonic stem cells into selective neurons on nanoscale ridge/groove pattern arrays. Biomaterials 2010, 31, 4360–4366. [Google Scholar] [CrossRef] [PubMed]
  52. McBeath, R.; Pirone, D.M.; Nelson, C.M.; Bhadriraju, K.; Chen, C.S. Cell shape, cytoskeletal tension, and RhoA regulate stem cell lineage commitment. Dev. Cell 2004, 6, 483–495. [Google Scholar] [CrossRef] [Green Version]
  53. Verkhratsky, A.; Ho, M.S.; Parpura, V. Evolution of Neuroglia. In Neuroglia in Neurodegenerative Diseases; Springer: New York, NY, USA, 2019; pp. 15–44. [Google Scholar]
  54. Georges, P.C.; Miller, W.J.; Meaney, D.F.; Sawyer, E.S.; Janmey, P.A. Matrices with compliance comparable to that of brain tissue select neuronal over glial growth in mixed cortical cultures. Biophys. J. 2006, 90, 3012–3018. [Google Scholar] [CrossRef] [Green Version]
  55. Doetsch, F. A niche for adult neural stem cells. Curr. Opin. Genet. Dev. 2003, 13, 543–550. [Google Scholar] [CrossRef]
  56. Alenghat, F.J.; Ingber, D.E. Mechanotransduction: All signals point to cytoskeleton, matrix, and integrins. Science’s STKE: Signal transduction knowledge environment. Sci. Signal. 2002, 119, pe6. [Google Scholar] [CrossRef]
  57. Kjellman, C.; Lidman, J.; Ljungström, K. Nilsson, Piezoelectric Sensor in a Living Organism for Fluid Pressure Measurement. U.S. Patent 6,886,411, 3 May 2005. [Google Scholar]
  58. Wada, Y.; Hayakawa, R. Piezoelectricity and pyroelectricity of polymers. Jpn. J. Appl. Phys. 1976, 15, 2041. [Google Scholar] [CrossRef]
  59. Parpaite, T.; Coste, B. Piezo channels. Curr. Biol. 2017, 27, R250–R252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Liu, Y.; Gao, J.; Peng, M.; Meng, H.; Ma, H.; Cai, P.; Si, G. A review on central nervous system effects of gastrodin. Front. Pharmacol. 2018, 9, 24. [Google Scholar] [CrossRef] [PubMed]
  61. Piccoli, A.; Rossettini, G.; Cecchetto, S.; Viceconti, A.; Ristori, D.; Turolla, A.; Testa, M. Effect of attentional focus instructions on motor learning and performance of patients with central nervous system and musculoskeletal disorders: A systematic review. J. Funct. Morphol. Kinesiol. 2018, 3, 40. [Google Scholar] [CrossRef] [Green Version]
  62. Schulte, F.; Kunin-Batson, A.S.; Olson-Bullis, B.A.; Banerjee, P.; Hocking, M.C.; Janzen, L.; Krull, K.R. Social attainment in survivors of pediatric central nervous system tumors: A systematic review and meta-analysis from the Children’s Oncology Group. J. Cancer Surviv. 2019, 13, 921–931. [Google Scholar] [CrossRef]
  63. Saheb, N.; Mekid, S. Fiber-Embedded Metallic Materials: From Sensing towards Nervous Behavior. Materials 2015, 8, 7938–7961. [Google Scholar] [CrossRef]
  64. Stavoe, A.K.; Holzbaur, E.L. Autophagy in Neurons. Annu. Rev. Cell Dev. Biol. 2019, 35, 477–500. [Google Scholar] [CrossRef]
  65. Swenarchuk, L.E. Nerve, Muscle, and Synaptogenesis. Cells 2019, 8, 1448. [Google Scholar] [CrossRef] [Green Version]
  66. Martí, D.; Brunel, N.; Ostojic, S. Correlations between synapses in pairs of neurons slow down dynamics in randomly connected neural networks. Phys. Rev. E 2018, 97, 062314. [Google Scholar] [CrossRef] [Green Version]
  67. Van Driesche, S.J.; Martin, K.C. New frontiers in RNA transport and local translation in neurons. Dev. Neurobiol. 2018, 78, 331–339. [Google Scholar] [CrossRef]
  68. Zhao, W.; Cui, W.; Xu, S.; Cheong, L.Z.; Wang, D.; Shen, C. Direct study of the electrical properties of PC12 cells and hippocampal neurons by EFM and KPFM. Nanoscale Adv. 2019, 1, 537–545. [Google Scholar] [CrossRef] [Green Version]
  69. Averbeck, B.B.; Lee, D. Coding and transmission of information by neural ensembles. Trends Neurosci. 2004, 27, 225–230. [Google Scholar] [CrossRef]
  70. Patel, N. Orientation of neurite growth by extracellar electric fields. J. Neurosci. 1982, 2, 483–496. [Google Scholar] [CrossRef]
  71. Freeman, J.A.; Manis, P.B.; Snipes, G.J. Steady growth cone currents revealed by a novel circularly vibrating probe: A possible mechanism underlying neurite growth. J. Neurosci. 1985, 13, 257–283. [Google Scholar] [CrossRef] [PubMed]
  72. Sisken, B.F.; Kanje, M.; Lundborg, G.; Herbst, E.; Kurtz, W. Stimulation of rat sciatic nerve regeneration with pulsed electromagnetic fields. Brain Res. 1989, 485, 309–316. [Google Scholar] [CrossRef]
  73. Kimura, K.; Yanagida, Y.; Haruyama, T.; Kobatake, E.; Aizawa, M. Gene expression in the electrically stimulated differentiation of PC12 cells. J. Biotechnol. 1998, 63, 55–65. [Google Scholar] [CrossRef]
  74. Kotwal, A.; Schmid, C.E. Electrical stimulation alters protein adsorption and nerve cell interactions with electrically conducting biomaterials. Biomaterials 2001, 22, 1055–1064. [Google Scholar] [CrossRef]
  75. Ghasemi-Mobarakeh, L.; Prabhakaran, M.P.; Morshed, M.; Nasr-Esfahani, M.H.; Baharvand, H.; Kiani, S.; Ramakrishna, S. Application of conductive polymers, scaffolds and electrical stimulation for nerve tissue engineering. J. Tissue Eng. Regen. Med. 2011, 5, 17–35. [Google Scholar] [CrossRef]
  76. Martin, R.M. Piezoelectricity. Phys. Rev. B 1972, 5, 1607. [Google Scholar] [CrossRef]
  77. Shamos, M.H.; Lavine, L.S. Piezoelectricity as a fundamental property of biological tissues. Nature 1967, 213, 267–269. [Google Scholar] [CrossRef]
  78. Telega, J.J.; Wojnar, R. Piezoelectric effects in biological tissues. J. Theor. Appl. Mech. 2002, 40, 723–759. [Google Scholar]
  79. Fukada, E. Electrical phenomena in biorheology. Biorheology 1982, 19, 15–27. [Google Scholar] [CrossRef] [PubMed]
  80. Athenstaedt, H. Pyroelectric and piezoelectric behaviour of human dental hard tissues. Arch. Oral Biol. 1971, 16, 495–501. [Google Scholar] [CrossRef]
  81. De Rossi, D.; Domenici, D.; Pastacaldi, P. Piezoelectric Properties of Dry Human Skin. IEEE Trans. Electr. Insul. 1985, 21, 511–517. [Google Scholar] [CrossRef]
  82. Ingber, D.E. The architecture of life. Sci. Am. 1998, 278, 48–57. [Google Scholar] [CrossRef]
  83. Reyes-Gasga, J.; Galindo-Mentle, M.; Brès, E.; Vargas-Becerril, N.; Orozco, E.; Rodríguez-Gómez, A.; García-García, R. Detection of the piezoelectricity effect in nanocrystals from human teeth. J. Phys. Chem. Solids 2020, 136, 109140. [Google Scholar] [CrossRef]
  84. Udovč, L.; Spreitzer, M.; Vukomanović, M. Towards hydrophilic piezoelectric poly-L-lactide films: Optimal processing, post-heat treatment and alkaline etching. Polym. J. 2019, 1–13. [Google Scholar] [CrossRef]
  85. Hoop, M.; Chen, X.; Ferrari, A.; Fajer, M.; Gagik, G.; Theo, T.; Dimos, P.; Bradley, N.; Salvador, P. Ultrasound-mediated piezoelectric differentiation of neuron-like PC12 cells on PVDF membranes. Sci. Rep. 2017, 7, 4028. [Google Scholar] [CrossRef] [Green Version]
  86. Ahn, A.C.; Grodzinsky, A.J. Relevance of collagen piezoelectricity to “Wolff’s Law”: A critical review. Med. Eng. Phys. 2009, 31, 733–741. [Google Scholar] [CrossRef] [Green Version]
  87. Lang, S.B. Pyroelectric Effect in Bone and Tendon. Nature 1966, 212, 704–705. [Google Scholar] [CrossRef]
  88. Anderson, J.C.; Eriksson, C. Electrical properties of wet collagen. Nature 1968, 218, 166–168. [Google Scholar] [CrossRef]
  89. Anderson, J.C.; Eriksson, C. Piezoelectric properties of dry and wet bone. Nature 1970, 227, 491–492. [Google Scholar] [CrossRef] [PubMed]
  90. Fukada, E. Piezoelectricity in polymers and biological materials. Ultrasonics 1968, 6, 229–234. [Google Scholar] [CrossRef]
  91. Furukawa, T. Piezoelectricity and pyroelectricity in polymers. IEEE Tran. Electr. Insul. 1989, 24, 375–394. [Google Scholar] [CrossRef]
  92. Puppi, D.; Chiellini, F.; Piras, A.; Chiellini, E. Polymeric materials for bone and cartilage repair. Prog. Polym. Sci. 2010, 35, 403–440. [Google Scholar] [CrossRef]
  93. Ribeiro, C.; Sencadas, V.; Correia, D.M.; Lanceros-Méndez, S. Piezoelectric polymers as biomaterials for tissue engineering applications. Colloids Surf. B Biointerfaces 2015, 136, 46–55. [Google Scholar] [CrossRef] [Green Version]
  94. Fukada, E. History and recent progress in piezoelectric polymers. IEEE Trans. Ultrason. Ferroelectr. Freq. Control 2000, 47, 1277–1290. [Google Scholar] [CrossRef]
  95. Zhong, Y.; Bellamkonda, R. V Biomaterials for the central nervous system. J. R. Soc. Interface 2008, 5, 957–975. [Google Scholar] [CrossRef]
  96. Jeznach, O.; Kołbuk, D.; Sajkiewicz, P. Injectable hydrogels and nanocomposite hydrogels for cartilage regeneration. J. Biomed. Mater. Res. A 2018, 106, 2762–2776. [Google Scholar] [CrossRef]
  97. Brown, B.N.; Badylak, S.F. Extracellular matrix as an inductive scaffold for functional tissue reconstruction. Transl. Res. 2014, 163, 268–285. [Google Scholar] [CrossRef] [Green Version]
  98. Ravichandran, R.; Astrand, C.; Patra, H.K.; Turner, A.P.; Chotteau, V.; Phopase, J. Intelligent ECM mimetic injectable scaffolds based on functional collagen building blocks for tissue engineering and biomedical applications. RSC Adv. 2017, 7, 21068–21078. [Google Scholar] [CrossRef] [Green Version]
  99. Miguel, S.P.; Sequeira, R.S.; Moreira, A.F.; Cabral, C.C.; Mendonça, A.G.; Ferreira, P.; Correia, I.J. An overview of electrospun membranes loaded with bioactive molecules for improving the wound healing process. Eur. J. Pharm. Biopharm. 2019, 139, 1–22. [Google Scholar] [CrossRef] [PubMed]
  100. Okamoto, M. The role of scaffolds in tissue engineering. In Handbook of Tissue Engineering Scaffolds; Elsevier: Amsterdam, The Netherlands, 2019; Volume 1, pp. 23–49. [Google Scholar]
  101. Morgado, P.I.; Aguiar-Ricardo, A.; Correia, I.J. Asymmetric membranes as ideal wound dressings: An overview on production methods, structure, properties and performance relationship. J. Membr. Sci. 2015, 490, 139–151. [Google Scholar] [CrossRef]
  102. Subramanian, A.; Krishnan, U.M.; Sethuraman, S. Development of biomaterial scaffold for nerve tissue engineering: Biomaterial mediated neural regeneration. J. Biomed. Sci. 2009, 16, 108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Shapiro, F. Overview of Deformities. In Pediatric Orthopedic Deformities; Springer: Cham, Switzerland, 2016; Volume 1, pp. 159–254. [Google Scholar]
  104. Royo-Gascon, N.; Wininger, M.; Scheinbeim, J.I.; Firestein, B.L.; Craelius, W. Piezoelectric substrates promote neurite growth in rat spinal cord neurons. Ann. Biomed. Eng. 2013, 41, 112–122. [Google Scholar] [CrossRef] [PubMed]
  105. Valentini, R.F.; Vargo, T.G.; Gardella, J.A., Jr.; Aebischer, P. Electrically charged polymeric substrates enhance nerve fibre outgrowth in vitro. Biomaterials 1992, 13, 183–190. [Google Scholar] [CrossRef]
  106. Aebischer, P.; Valentini, R.F.; Dario, P.; Domenici, C.; Galletti, P.M. Piezoelectric guidance channels enhance regeneration in the mouse sciatic nerve after axotomy. Brain Res. 1987, 436, 165–168. [Google Scholar] [CrossRef]
  107. Delaviz, H.; Faghihi, A.; Delshad, A.A.; Hadi Bahadori, M.; Mohamadi, J.; Roozbehi, A. Repair of peripheral nerve defects using a polyvinylidene fluoride channel containing nerve growth factor and collagen gel in adult rats. Cell J. 2011, 13, 137–142. [Google Scholar]
  108. Young, T.H.; Chang, H.H.; Lin, D.J.; Cheng, L.P. Surface modification of microporous PVDF membranes for neuron culture. J. Membr. Sci. 2010, 350, 32–41. [Google Scholar] [CrossRef]
  109. Ariga, K.; Jia, X.; Song, J.; Hsieh, C.T.; Hsu, S.H. Materials Nanoarchitectonics as Cell Regulators. ChemNanoMat 2019, 5, 692–702. [Google Scholar] [CrossRef]
  110. Ai, J.; Kiasat-Dolatabadi, A.; Ebrahimi-Barough, S.; Ai, A.; Lotfibakhshaiesh, N.; Norouzi-Javidan, A.; Aghayan, H.R. Polymeric scaffolds in neural tissue engineering: A review. Arch Neurosci. 2014, 1, 15–20. [Google Scholar] [CrossRef] [Green Version]
  111. Abzan, N.; Kharaziha, M.; Labbaf, S. Development of three-dimensional piezoelectric polyvinylidene fluoride-graphene oxide scaffold by non-solvent induced phase separation method for nerve tissue engineering. Mater. Design 2019, 167, 107636. [Google Scholar] [CrossRef]
  112. Khorshidi, S.; Ansari, S.; Naghizadeh, Z.; Akbari, N.; Karkhaneh, A.; Haghighipour, N. Concurrent effects of piezoelectricity and hydrostatic pressure on chondrogenic differentiation of stem cells. Mater. Lett. 2019, 246, 71–75. [Google Scholar] [CrossRef]
  113. Lee, Y.S.; Arinzeh, T.L. The influence of piezoelectric scaffolds on neural differentiation of human neural stem/progenitor cells. Tissue Eng. A 2012, 18, 2063–2072. [Google Scholar] [CrossRef] [PubMed]
  114. Lee, Y.S.; Collins, G.; Arinzeh, T.L. Neurite extension of primary neurons on electrospun piezoelectric scaffolds. Acta Biomater. 2011, 7, 3877–3886. [Google Scholar] [CrossRef]
  115. Genchi, G.G.; Sinibaldi, E.; Ceseracciu, L.; Labardi, M.; Marino, A.; Marras, S.; Ciofani, G. Ultrasound-activated piezoelectric P (VDF-TrFE)/boron nitride nanotube composite films promote differentiation of human SaOS-2 osteoblast-like cells. Nanomedicine 2018, 14, 2421–2432. [Google Scholar] [CrossRef]
  116. Fine, E.G.; Valentini, R.F.; Bellamkonda, R.; Aebischer, P. Improved nerve regeneration through piezoelectric vinylidenefluoride-trifluoroethylene copolymer guidance channels. Biomaterials 1991, 12, 775–780. [Google Scholar] [CrossRef]
  117. Wang, A.; Hu, M.; Zhou, L.; Qiang, X. Self-Powered Well-Aligned P (VDF-TrFE) Piezoelectric Nanofiber Nanogenerator for Modulating an Exact Electrical Stimulation and Enhancing the Proliferation of Preosteoblasts. Nanomaterials 2019, 9, 349. [Google Scholar] [CrossRef] [Green Version]
  118. Collazos-Castro, J.E.; Polo, J.L.; Hernández-Labrado, G.R.; Padial-Cañete, V.; García-Rama, C. Bioelectrochemical control of neural cell development on conducting polymers. Biomaterials 2010, 31, 9244–9255. [Google Scholar] [CrossRef]
  119. Ludwig, K.A.; Uram, J.D.; Yang, J.; Martin, D.C.; Kipke, D.R. Chronic neural recordings using silicon microelectrode arrays electrochemically deposited with a poly (3, 4-ethylenedioxythiophene)(PEDOT) film. J. Neural. Eng. 2006, 3, 59–70. [Google Scholar] [CrossRef] [Green Version]
  120. Pires, F.; Ferreira, Q.; Rodrigues, C.A.; Morgado, J.; Ferreira, F.C. Neural stem cell differentiation by electrical stimulation using a cross-linked PEDOT substrate: Expanding the use of biocompatible conjugated conductive polymers for neural tissue engineering. Biochim. Biophys. Acta 2015, 1850, 1158–1168. [Google Scholar] [CrossRef]
  121. Sebaa, M.; Nguyen, T.Y.; Dhillon, S.; Garcia, S.; Liu, H. The effects of poly (3, 4-ethylenedioxythiophene) coating on magnesium degradation and cytocompatibility with human embryonic stem cells for potential neural applications. J. Biomed. Mater. Res. A 2015, 103, 25–37. [Google Scholar] [CrossRef]
  122. Du, L.; Li, T.; Jin, F.; Wang, Y.; Li, R.; Zheng, J.; Feng, Z.Q. Design of high conductive and piezoelectric poly (3, 4-ethylenedioxythiophene)/chitosan nanofibers for enhancing cellular electrical stimulation. J. Colloid Interface Sci. 2020, 559, 65–75. [Google Scholar] [CrossRef] [PubMed]
  123. Evans, G.R.; Brandt, K.; Niederbichler, A.D.; Chauvin, P.; Hermann, S.; Bogle, M.; Patrick, C.W. Clinical long-term in vivo evaluation of poly (L-lactic acid) porous conduits for peripheral nerve regeneration. J. Biomater. Sci. Polym. 2000, 11, 869–878. [Google Scholar] [CrossRef] [PubMed]
  124. Jia, L.; Prabhakaran, M.P.; Qin, X.; Ramakrishna, S. Stem cell differentiation on electrospun nanofibrous substrates for vascular tissue engineering. Mater. Sci. Eng. C 2013, 33, 4640–4650. [Google Scholar] [CrossRef] [PubMed]
  125. Yang, F.; Murugan, R.; Ramakrishna, S.; Wang, X.; Ma, Y.X.; Wang, S. Fabrication of nano-structured porous PLLA scaffold intended for nerve tissue engineering. Biomaterials 2004, 25, 1891–1900. [Google Scholar] [CrossRef]
  126. Yang, F.; Murugan, R.; Wang, S.; Ramakrishna, S. Electrospinning of nano/micro scale poly (L-lactic acid) aligned fibers and their potential in neural tissue engineering. Biomaterials 2005, 26, 2603–2610. [Google Scholar] [CrossRef]
  127. Prabhakaran, M.P.; Venugopal, J.; Ramakrishna, S. Electrospun nanostructured scaffolds for bone tissue engineering. Acta Biomater. 2009, 5, 2884–2893. [Google Scholar] [CrossRef]
  128. Prabhakaran, M.P.; Ghasemi-Mobarakeh, L.; Jin, G.; Ramakrishna, S. Electrospun conducting polymer nanofibers and electrical stimulation of nerve stem cells. J. Biosci. Bioeng. 2011, 112, 501–507. [Google Scholar] [CrossRef]
  129. Jacob, J.; More, N.; Mounika, C.; Gondaliya, P.; Kalia, K.; Kapusetti, G. The Smart Piezoelectric Nanohybrid of Poly-(3-hydroxybutyrate-co-3-hydroxyvalerate) and Barium Titanate for Stimulated Cartilage Regeneration. ACS Appl. Bio Mater. 2019, 2, 4922–4931. [Google Scholar] [CrossRef]
  130. De Guzman, R.C.; Loeb, J.A.; VandeVord, P.J. Electrospinning of matrigel to deposit a basal lamina-like nanofiber surface. J. Biomater. Sci. Polym. Ed. 2010, 21, 1081–1101. [Google Scholar] [CrossRef]
  131. O’Shaughnessy, T.J.; Lin, H.J.; Ma, W. Functional synapse formation among rat cortical neurons grown on three-dimensional collagen gels. Neurosci. Lett. 2003, 340, 169–172. [Google Scholar] [CrossRef]
  132. Shuai, C.; Liu, G.; Yang, Y.; Yang, W.; He, C.; Wang, G.; Peng, S. Functionalized BaTiO3 enhances piezoelectric effect towards cell response of bone scaffold. Colloids Surf. B. 2020, 185, 110587. [Google Scholar] [CrossRef] [PubMed]
  133. Mercadelli, E.; Sanson, A.; Galassi, C. Porous Piezoelectric Ceramics. In Piezoelectric Ceramics; Suaste-Gomez, E., Ed.; InTech: Rijeka, Croatia, 2010; pp. 111–129. [Google Scholar]
  134. Wersing, W.; Lubitz, K.; Mohaupt, J. Dielectric, elastic and piezoelectric properties of porous PZT ceramics. Ferroelectrics 1986, 68, 77–97. [Google Scholar] [CrossRef]
  135. Ringgaard, E.; Lautzenhiser, F.; Bierregaard, L.; Zawada, T.; Molz, E. Development of porous piezoceramics for medical and sensor applications. Materials 2015, 8, 8877–8889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Xue, C.; Hu, N.; Gu, Y.; Yang, Y.; Liu, Y.; Liu, J.; Ding, F.; Gu, X. Joint Use of a Chitosan/PLGA Scaffold and MSCs to Bridge an Extra Large Gap in Dog Sciatic Nerve. Neurorehabil. Neural Repair 2012, 26, 96–106. [Google Scholar] [CrossRef]
  137. Carville, N.C.; Collins, L.; Manzo, M.; Gallo, K.; Lukasz, B.I.; McKayed, K.K.; Rodriguez, B.J. Biocompatibility of ferroelectric lithium niobate and the influence of polarization charge on osteoblast proliferation and function. J. Biomed. Mater. Res. A 2015, 103, 2540–2548. [Google Scholar] [CrossRef]
  138. Furuya, K.; Morita, Y.; Tanaka, K.; Katayama, T.; Nakamachi, E. Acceleration of osteogenesis by using barium titanate piezoelectric ceramic as an implant material. In Proceedings of the International Society for Optics and Photonics, San Diego, CA, USA, 3–7 May 2011; Volume 7975, p. 79750U. [Google Scholar]
  139. Ball, J.P.; Mound, B.A.; Nino, J.C.; Allen, J.B. Biocompatible evaluation of barium titanate foamed ceramic structures for orthopedic applications. J. Biomed. Mater. Res. A 2014, 102, 2089–2095. [Google Scholar] [CrossRef]
  140. Lopes, H.B.; Santos, T.D.S.; De Oliveira, F.S.; Freitas, G.P.; De Almeida, A.L.; Gimenes, R.; Beloti, M.M. Poly (vinylidene-trifluoroethylene)/barium titanate composite for in vivo support of bone formation. J. Biomater. Appl. 2014, 29, 104–112. [Google Scholar] [CrossRef]
  141. Zhang, X.; Zhang, C.; Lin, Y.; Hu, P.; Shen, Y.; Wang, K.; Liu, Y. Nanocomposite membranes enhance bone regeneration through restoring physiological electric microenvironment. ACS Nano 2016, 10, 7279–7286. [Google Scholar] [CrossRef]
  142. Roberts, S. Dielectric and piezoelectric properties of barium titanate. Phys. Rev. 1947, 71, 890. [Google Scholar] [CrossRef]
  143. Baxter, F.R.; Bowen, C.R.; Turner, I.G.; Dent, A.C. Electrically active bioceramics: A review of interfacial responses. Ann. Biomed. Eng. 2010, 38, 2079–2092. [Google Scholar] [CrossRef] [PubMed]
  144. Ciofani, G.; Ricotti, L.; Canale, C.; D’Alessandro, D.; Berrettini, S.; Mazzolai, B.; Mattoli, V. Effects of barium titanate nanoparticles on proliferation and differentiation of rat mesenchymal stem cells. Colloids Surf. B Biointerfaces 2013, 102, 312–320. [Google Scholar] [CrossRef]
  145. Ciofani, G.; Ricotti, L.; Mattoli, V. Preparation, characterization and in vitro testing of poly (lactic-co-glycolic) acid/barium titanate nanoparticle composites for enhanced cellular proliferation. Biomed. Microdevices 2011, 13, 255–266. [Google Scholar] [CrossRef] [PubMed]
  146. Ivanova, O.; Williams, C.; Campbell, T. Additive manufacturing (AM) and nanotechnology: Promises and challenges. Rapid Prototyp. J. 2013, 19, 353–364. [Google Scholar] [CrossRef] [Green Version]
  147. Jacob, J.; More, N.; Kalia, K.; Kapusetti, G. Piezoelectric smart biomaterials for bone and cartilage tissue engineering. Inflamm. Regen. 2018, 38, 2. [Google Scholar] [CrossRef] [Green Version]
  148. Matassi, F.; Nistri, L.; Paez, D.C.; Innocenti, M. New biomaterials for bone regeneration. Clin. Cases Miner. Bone Metab. 2011, 8, 21. [Google Scholar]
  149. Ahmad, P.; Khandaker, M.U.; Khan, Z.R.; Amin, Y.M. Synthesis of boron nitride nanotubes via chemical vapour deposition: A comprehensive review. Mater. Sci. Eng. R. 2010, 70, 92–111. [Google Scholar] [CrossRef]
  150. Lahiri, D.; Rouzaud, F.; Richard, T.; Keshri, A.K.; Bakshi, S.R.; Kos, L.; Agarwal, A. Boron nitride nanotube reinforced polylactide–polycaprolactone copolymer composite: Mechanical properties and cytocompatibility with osteoblasts and macrophages in vitro. Acta Biomater. 2010, 6, 3524–3533. [Google Scholar] [CrossRef]
  151. Lahiri, D.; Singh, V.; Benaduce, A.P.; Seal, S.; Kos, L.; Agarwal, A. Boron nitride nanotube reinforced hydroxyapatite composite: Mechanical and tribological performance and in-vitro biocompatibility to osteoblasts. J. Mech. Behav. Biomed. 2011, 4, 44–56. [Google Scholar] [CrossRef]
  152. Li, X.; Zhi, C.; Hanagata, N.; Yamaguchi, M.; Bando, Y.; Golberg, D. Boron nitride nanotubes functionalized with mesoporous silica for intracellular delivery of chemotherapy drugs. Chem. Commun. 2013, 49, 7337–7339. [Google Scholar] [CrossRef]
  153. Ciofani, G.; Danti, S.; Genchi, G.G.; Mazzolai, B.; Mattoli, V. Boron nitride nanotubes: Biocompatibility and potential spill-over in nanomedicine. Small 2013, 9, 1672–1685. [Google Scholar] [CrossRef] [PubMed]
  154. Li, X.; Hanagata, N.; Wang, X.; Yamaguchi, M.; Yi, W.; Bando, Y.; Golberg, D. Multimodal luminescent-magnetic boron nitride nanotubes@ NaGdF 4: Eu structures for cancer therapy. Chem. Commun. 2014, 50, 4371–4374. [Google Scholar] [CrossRef]
  155. Weng, Q.; Wang, B.; Wang, X.; Hanagata, N.; Li, X.; Liu, D.; Golberg, D. Highly water-soluble, porous, and biocompatible boron nitrides for anticancer drug delivery. ACS Nano 2014, 8, 6123–6130. [Google Scholar] [CrossRef] [PubMed]
  156. Wang, J.; Lee, C.H.; Yap, Y.K. Recent advancements in boron nitride nanotubes. Nanoscale 2010, 2, 2028–2034. [Google Scholar] [CrossRef] [PubMed]
  157. Ciofani, G.; Raffa, V.; Menciassi, A.; Cuschieri, A. Boron nitride nanotubes: An innovative tool for nanomedicine. Nano Today 2009, 4, 8–10. [Google Scholar] [CrossRef]
  158. Rasmussen, J.W.; Martinez, E.; Louka, P.; Wingett, D.G. Zinc oxide nanoparticles for selective destruction of tumor cells and potential for drug delivery applications. Expert Opin. Drug Deliv. 2010, 7, 1063–1077. [Google Scholar] [CrossRef] [Green Version]
  159. Goel, S.; Kumar, B. A review on piezo-/ferro-electric properties of morphologically diverse ZnO nanostructures. J. Alloys Compd. 2019, 816, 152491. [Google Scholar] [CrossRef]
  160. Yin, Y.; Lin, Q.; Sun, H.; Chen, D.; Wu, Q.; Chen, X.; Li, S. Cytotoxic effects of ZnO hierarchical architectures on RSC96 Schwann cells. Res. Lett. 2012, 7, 439. [Google Scholar] [CrossRef] [Green Version]
  161. Safaei, M.; Sodano, H.A.; Anton, S.R. A review of energy harvesting using piezoelectric materials: State-of-the-art a decade later (2008–2018). Smart Mater. Struct. 2019, 28, 113001. [Google Scholar] [CrossRef]
  162. Ribeiro, C.; Correia, D.M.; Ribeiro, S.; Sencadas, V.; Botelho, G.; Lanceros-Méndez, S. Piezoelectric poly (vinylidene fluoride) microstructure and poling state in active tissue engineering. Eng. Life Sci. 2015, 15, 351–356. [Google Scholar] [CrossRef] [Green Version]
  163. Aguilar, M.R.; San Román, J. Introduction to smart polymers and their applications. In Smart Polymers and Their Applications; Woodhead Publishing: Sawston, UK, 2019; pp. 1–11. [Google Scholar]
  164. Piskin, E. Biodegradable polymers as biomaterials. J. Biomater. Sci. Polym. Ed. 1995, 6, 775–795. [Google Scholar] [CrossRef] [PubMed]
  165. Sajkiewicz, P. Crystallization behaviour of poly(vinylidene fluoride). Eur. Polym. J. 1999, 35, 1581–1590. [Google Scholar] [CrossRef]
  166. Gradys, A.; Sajkiewicz, P.; Adamovsky, S.; Minakov, A.A.; Schick, C. Crystallization of poly(vinylidene fluoride) during ultra-fast cooling. Thermochim. Acta 2007, 461, 153–157. [Google Scholar] [CrossRef]
  167. Esterly, D.M.; Love, B.J. Phase transformation to β-poly (vinylidene fluoride) by milling. J. Polym. Sci. B Polym. Phys. 2004, 42, 91–97. [Google Scholar] [CrossRef]
  168. Cozza, E.S.; Monticelli, O.; Marsano, E.; Cebe, S. On the Electrospinning of PVDF: Influence of the Experimental Conditions on the Nanofiber Properties. Polym. Int. 2013, 62, 41–48. [Google Scholar] [CrossRef]
  169. Yu, L.; Cebe, P. Crystal polymorphism in electrospun composite nanofibers of poly (vinylidene fluoride) with nanoclay. Polymer 2009, 50, 2133–2141. [Google Scholar] [CrossRef]
  170. El Mohajir, B.E.; Heymans, N. Changes in structural and mechanical behaviour of PVDF with processing and thermomechanical treatments. 1. Change in structure. Polymer 2001, 42, 5661–5667. [Google Scholar] [CrossRef]
  171. Imamura, R.; Silva, A.B.; Gregorio, R., Jr. γ→ β Phase transformation induced in poly (vinylidene fluoride) by stretching. J. Appl. Polym. Sci. 2008, 110, 3242–3246. [Google Scholar] [CrossRef]
  172. Wang, J.; Li, H.; Liu, J.; Duan, Y.; Jiang, S.; Yan, Y. On the α→β Transition of Carbon Coated Highly Oriented PVDF Ultrathin Film Induced by Melt Recrystallization. J. Am. Chem. Soc. 2003, 125, 1496–1497. [Google Scholar] [CrossRef]
  173. Kaura, T.; Nath, R.; Perlman, M.M. Simultaneous stretching and corona poling of PVDF films. J. Phys. D: Appl. Phys. 1991, 24, 1848. [Google Scholar] [CrossRef]
  174. Ramanathan, A.K.; Headings, L.M.; Dapino, M.J. Design optimization of flexible piezoelectric PVDF unimorphs for surface pressure transducer applications. In Smart Structures and NDE for Energy Systems and Industry 4.0; International Society for Optics and Photonics: Bellingham, WA, USA, 2019; Volume 10973, p. 1097307. [Google Scholar]
  175. Ellingford, C.; Smith, H.; Yan, X.; Bowen, C.; Figiel, Ł.; McNally, T.; Wan, C. Electrical dual-percolation in MWCNTs/SBS/PVDF based thermoplastic elastomer (TPE) composites and the effect of mechanical stretching. Eur. Polym. J. 2019, 112, 504–514. [Google Scholar] [CrossRef]
  176. Zhang, S.; Jia, Z.; Liu, T.; Wei, G.; Su, Z. Electrospinning Nanoparticles-Based Materials Interfaces for Sensor Applications. Sensors 2019, 19, 3977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Li, Y.; Liao, C.; Tjong, S.C. Electrospun Polyvinylidene Fluoride-Based Fibrous Scaffolds with Piezoelectric Characteristics for Bone and Neural Tissue Engineering. Nanomaterials 2019, 9, 952. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Yu, L.; Zhou, P.; Wu, D.; Wang, L.; Lin, L.; Sun, D. Shoepad nanogenerator based on electrospun PVDF nanofibers. Microsyst. Technol. 2019, 25, 3151–3156. [Google Scholar] [CrossRef]
  179. Ribeiro, C.; Costa, C.M.; Correia, D.M.; Nunes-Pereira, J.; Oliveira, J.; Martins, P.; Lanceros-Méndez, S. Electroactive poly (vinylidene fluoride)-based structures for advanced applications. Nat. Protoc. 2018, 13, 681. [Google Scholar] [CrossRef] [PubMed]
  180. Fortunato, M.; Cavallini, D.; De Bellis, G.; Marra, F.; Tamburrano, A.; Sarto, F.; Sarto, M.S. Phase Inversion in PVDF Films with Enhanced Piezoresponse Through Spin-Coating and Quenching. Polymers 2019, 11, 1096. [Google Scholar] [CrossRef] [Green Version]
  181. Liu, Z.H.; Pan, C.T.; Lin, L.W.; Huang, J.C.; Ou, Z.Y. Direct-write PVDF nonwoven fiber fabric energy harvesters via the hollow cylindrical near-field electrospinning process. Smart Mater. Struct. 2013, 23, 025003. [Google Scholar] [CrossRef]
  182. Zaarour, B.; Zhu, L.; Jin, X. Controlling the surface structure, mechanical properties, crystallinity, and piezoelectric properties of electrospun PVDF nanofibers by maneuvering molecular weight. Soft Mater. 2019, 17, 181–189. [Google Scholar] [CrossRef]
  183. Singh, R.K.; Lye, S.W.; Miao, J. PVDF Nanofiber Sensor for Vibration Measurement in a String. Sensors 2019, 19, 3739. [Google Scholar] [CrossRef] [Green Version]
  184. Khalifa, M.; Janakiraman, S.; Ghosh, S.; Venimadhav, A.; Anandhan, S. PVDF/halloysite nanocomposite-based non-wovens as gel polymer electrolyte for high safety lithium ion battery. Polym. Compos. 2019, 40, 2320–2334. [Google Scholar] [CrossRef]
  185. Liang, S.; Kang, Y.; Tiraferri, A.; Giannelis, E.P.; Huang, X.; Elimelech, M. Highly hydrophilic polyvinylidene fluoride (PVDF) ultrafiltration membranes via postfabrication grafting of surface-tailored silica nanoparticles. ACS Appl. Mater. Interfaces 2013, 5, 6694–6703. [Google Scholar] [CrossRef] [PubMed]
  186. Mandal, D.; Henkel, K.; Schmeißer, D. The electroactive β-phase formation in poly (vinylidene fluoride) by gold nanoparticles doping. Materials Lett. 2012, 73, 123–125. [Google Scholar] [CrossRef]
  187. Li, J.H.; Shao, X.S.; Zhou, Q.; Li, M.Z.; Zhang, Q.Q. The double effects of silver nanoparticles on the PVDF membrane: Surface hydrophilicity and antifouling performance. App. Surf. Sci. 2013, 265, 663–670. [Google Scholar] [CrossRef]
  188. Liang, S.; Xiao, K.; Mo, Y.; Huang, X. A novel ZnO nanoparticle blended polyvinylidene fluoride membrane for anti-irreversible fouling. J. Membr. Sci. 2012, 394, 184–192. [Google Scholar] [CrossRef]
  189. Teow, Y.H.; Ahmad, A.L.; Lim, J.K.; Ooi, B.S. Preparation and characterization of PVDF/TiO2 mixed matrix membrane via in situ colloidal precipitation method. Desalination 2012, 295, 61–69. [Google Scholar] [CrossRef]
  190. Zhang, J.; Xu, Z.; Mai, W.; Min, C.; Zhou, B.; Shan, M.; Qian, X. Improved hydrophilicity, permeability, antifouling and mechanical performance of PVDF composite ultrafiltration membranes tailored by oxidized low-dimensional carbon nanomaterials. J. Mater. Chem. A 2013, 1, 3101–3111. [Google Scholar] [CrossRef]
  191. Song, H.; Shao, J.; He, Y.; Liu, B.; Zhong, X. Natural organic matter removal and flux decline with PEG–TiO2-doped PVDF membranes by integration of ultrafiltration with photocatalysis. J. Membr. Sci. 2012, 405, 48–56. [Google Scholar] [CrossRef]
  192. Li, N.; Xiao, C.; An, S.; Hu, X. Preparation and properties of PVDF/PVA hollow fiber membranes. Desalination 2010, 250, 530–537. [Google Scholar] [CrossRef]
  193. Gayen, A.L.; Mondal, D.; Roy, D.; Bandyopadhyay, P.; Manna, S.; Basu, R.; Nandy, P. Improvisation of electrical properties of PVDF-HFP: Use of novel metallic nanoparticles. J. Mater. Sci.: Mater. 2017, 28, 14798–14808. [Google Scholar] [CrossRef]
  194. Jaleh, B.; Sodagar, S.; Momeni, A.; Jabbari, A. Nanodiamond particles/PVDF nanocomposite flexible films: Thermal, mechanical and physical properties. Mater. Res. Express 2016, 3, 085028. [Google Scholar] [CrossRef]
  195. Fraczek-Szczypta, A. Carbon nanomaterials for nerve tissue stimulation and regeneration. Mater. Sci. Eng. C 2014, 34, 35–49. [Google Scholar] [CrossRef]
  196. Tsonos, C.; Pandis, C.; Soin, N.; Sakellari, D.; Myrovali, E.; Kripotou, S.; Siores, E. Multifunctional nanocomposites of poly (vinylidene fluoride) reinforced by carbon nanotubes and magnetite nanoparticles. Express Polym. Lett. 2015, 9. [Google Scholar] [CrossRef]
  197. Lorber, B.; Hsiao, W.K.; Hutchings, I.M.; Martin, K.R. Adult rat retinal ganglion cells and glia can be printed by piezoelectric inkjet printing. Biofabrication 2013, 6, 015001. [Google Scholar] [CrossRef] [PubMed]
  198. Inaoka, T.; Shintaku, H.; Nakagawa, T.; Kawano, S.; Ogita, H.; Sakamoto, T.; Ito, J. Piezoelectric materials mimic the function of the cochlear sensory epithelium. Proc. Natl. Acad. Sci. USA 2011, 108, 18390–18395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Gao, Y.; Wang, Z.L. Electrostatic potential in a bent piezoelectric nanowire. The fundamental theory of nanogenerator and nanopiezotronics. Nano Lett. 2007, 7, 2499–2505. [Google Scholar] [CrossRef] [PubMed]
  200. Walsh, J.F.; Manwaring, M.E.; Tresco, P.A. Directional neurite outgrowth is enhanced by engineered meningeal cell-coated substrates. Tissue Eng. 2005, 11, 1085–1094. [Google Scholar] [CrossRef] [PubMed]
  201. Asano, T.; Kubo, T.; Nishikitani, Y. Electrochemical properties of dye-sensitized solar cells fabricated with PVDF-type polymeric solid electrolytes. J. Photochem. Photobiol. 2004, 164, 111–115. [Google Scholar] [CrossRef]
  202. Marino, A.; Arai, S.; Hou, Y.; Sinibaldi, E.; Pellegrino, M.; Chang, B.; Mazzolai, V.; Mattoli Suzuki, M.; Ciofani, G. Piezoelectric nanoparticle-assisted wireless neuronal stimulation. ACS Nano 2015, 9, 7678–7689. [Google Scholar] [CrossRef]
  203. Genchi, G.G.; Ceseracciu, L.; Marino, A.; Labardi, M.; Marras, S.; Pignatelli, F.; Bruschini, L.; Mattoli, V.; Ciofani, G. P(VDF-TrFE)/BaTiO3 nanoparticle composite films mediate piezoelectric stimulation and promote differentiation of SH-SY5Y neuroblastoma cells. Adv. Healthc. Mater. 2016, 5, 1808–1820. [Google Scholar] [CrossRef]
  204. Weber, N.; Lee, Y.S.; Shanmugasundaram, S.; Jaffe, M.; Arinzeh, T.L. Characterization and in vitro cytocompatibility of piezoelectric electrospun scaffolds. Acta Biomater. 2010, 6, 3550–3556. [Google Scholar] [CrossRef]
  205. De Ruiter, G.C.; Malessy, M.J.; Yaszemski, M.J.; Windebank, A.J.; Spinner, R.J. Designing ideal conduits for peripheral nerve repair. Neurosurg. Focus. 2009, 26, E5. [Google Scholar] [CrossRef] [PubMed]
  206. Ichihara, S.; Inada, Y.; Nakamura, T. Artificial nerve tubes and their application for repair of peripheral nerve injury: An update of current concepts. Injury 2008, 39, 29–39. [Google Scholar] [CrossRef] [PubMed]
  207. Ji, Y.; Jin, R.; Zhang, X.; Bouchilaoun, R.; Fan, J.; Zhao, R.; Yang, H. Electric polarizations in PVDF-TrFE nanorods under lateral nanoshaping. Int. J. Appl. 2019, 126, 174108. [Google Scholar] [CrossRef]
  208. Wan, C.; Bowen, C.R. Multiscale-structuring of polyvinylidene fluoride for energy harvesting: The impact of molecular-, micro- and macro-structure. J. Mater. Chem. A 2017, 5, 3091–3128. [Google Scholar] [CrossRef] [Green Version]
  209. Sun, F.C.; Dongare, A.M.; Asandei, A.D.; Alpay, S.P.; Nakhmanson, S. Temperature dependent structural, elastic, and polar properties of ferroelectric polyvinylidene fluoride (PVDF) and trifluoroethylene (TrFE). copolymers. J. Mater. Chem. C 2015, 3, 8389–8396. [Google Scholar] [CrossRef]
  210. Belkas, J.S.; Shoichet, M.S.; Midha, R. Peripheral nerve regeneration through guidance tubes. Neurol. Res. 2004, 26, 151–160. [Google Scholar] [CrossRef]
  211. Boni, R.; Ali, A.; Shavandi, A.; Clarkson, A.N. Current and novel polymeric biomaterials for neural tissue engineering. J Biomed. Sci. 2018, 25, 90. [Google Scholar] [CrossRef] [Green Version]
  212. Fine, E.G.; Valentini, R.F.; Bellamkonda, R.; Aebischer, P. Influence of surface texture of polymeric sheets throught piezoelectric vinylidenefluoride-trifluoroethylene copolymer guidance channels. Biomaterials 1991, 12, 259–263. [Google Scholar] [CrossRef]
  213. Martins, P.M.; Ribeiro, S.; Ribeiro, C.; Sencadas, V.; Gomes, A.C.; Gama, F.M.; Lanceros-Méndez, S. Effect of poling state and morphology of piezoelectric poly(vinylidene fluoride) membranes for skeletal muscle tissue engineering. RSC Adv. 2013, 3, 17938–17944. [Google Scholar] [CrossRef] [Green Version]
  214. Ke, S.; Huang, H.; Ren, L.; Wang, Y. Nearly constant dielectric loss behavior in poly (3-hydroxybutyrate-co-3-hydroxyvalerate) biodegradable polyester. J. Appl. Phys. 2009. [Google Scholar] [CrossRef] [Green Version]
  215. Numata, K.; Abe, H.; Doi, Y. Enzymatic processes for biodegradation of poly (hydroxyalkanoate) s crystals. Can. J. Chem. 2008, 86, 471–483. [Google Scholar] [CrossRef]
  216. Willerth, S.M.; Sakiyama-Elbert, S.E. Approaches to neural tissue engineering using scaffolds for drug delivery. Adv. Drug Deliv. Rev. 2007, 59, 325–338. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Wu, Q.; Wang, Y.; Chen, G.Q. Medical application of microbial biopolyesters polyhydroxyalkanoates. Artif. Cells Blood Substit. Immobil. Biotechnol. 2009, 37, 1–12. [Google Scholar] [CrossRef] [PubMed]
  218. Misra, S.K.; Valappil, S.P.; Roy, I.; Boccaccini, A.R. Polyhydroxyalkanoate (PHA)/inorganic phase composites for tissue engineering applications. Biomacromolecules 2006, 7, 2249–2258. [Google Scholar] [CrossRef]
  219. Prabhakaran, M.P.; Vatankhah, E.; Ramakrishna, S. Electrospun aligned PHBV/collagen nanofibers as substrates for nerve tissue engineering. Biotechnol Bioeng. 2013, 110, 2775–2784. [Google Scholar] [CrossRef]
  220. Rahman, M.S.; Tsuchiya, T. Enhancement of chondrogenic differentiation of human articular chondrocytes by biodegradable polymers. Tissue Eng. 2001, 7, 781–790. [Google Scholar] [CrossRef]
  221. Chen, W.; Tong, Y.W. PHBV microspheres as neural tissue engineering scaffold support neuronal cell growth and axon–dendrite polarization. Acta Biomater. 2012, 8, 540–548. [Google Scholar] [CrossRef]
  222. Rivera-Briso, A.L.; Serrano-Aroca, A. Poly(3-Hydroxybutyrate-co-3-Hydroxyvalerate): Enhancement Strategies for Advanced Applications. Polymers 2018, 10, 732. [Google Scholar] [CrossRef] [Green Version]
  223. Sencadas, V.; Ribeiro, C.; Heredia, A.; Bdikin, I.K.; Kholkin, A.L.; Lanceros-Méndez, S. Local piezoelectric activity of single poly (L-lactic acid)(PLLA) microfibers. Appl. Phys. A 2012, 109, 51–55. [Google Scholar] [CrossRef]
  224. Jin, L.; Feng, Z.Q.; Zhu, M.L.; Wang, T.; Leach, M.K.; Jiang, Q. A novel fluffy conductive polypyrrole nano-layer coated PLLA fibrous scaffold for nerve tissue engineering. J. Biomed. Nanotechnol. 2012, 8, 779–785. [Google Scholar] [CrossRef]
  225. Zhang, K.; Zheng, H.; Liang, S.; Gao, C. Aligned PLLA nanofibrous scaffolds coated with graphene oxide for promoting neural cell growth. Acta Biomater. 2016, 37, 131–142. [Google Scholar] [CrossRef] [PubMed]
  226. Zuidema, J.M.; Provenza, C.; Caliendo, T.; Dutz, S.; Gilbert, R.J. Magnetic NGF-releasing PLLA/iron oxide nanoparticles direct extending neurites and preferentially guide neurites along aligned electrospun microfibers. ACS Chem. Neurosci. 2015, 6, 1781–1788. [Google Scholar] [CrossRef] [PubMed]
  227. Venugopal, J.; Zhang, Y.Z.; Ramakrishna, S. Electrospun nanofibres: Biomedical applications. Proceedings of the institution of mechanical engineers. N J. Nanoeng. Nanosyst. 2004, 218, 35–45. [Google Scholar]
  228. Philipp, B.; Bock, W.; Schierbaum, F. Application of polysaccharides and their derivatives as supporting materials and auxiliary substances in medicine and nutrition. J. Polym. Sci. Polym. Symp. 1979, 66, 83–100. [Google Scholar] [CrossRef]
  229. Franz, G. Polysaccharides in pharmacy. Adv. Polym. Sci. 1986, 76, 1–30. [Google Scholar]
  230. Miyamoto, T.; Takahashi, S.I.; Ito, H.; Inagaki, H.; Noishiki, Y. Tissue biocompatibility of cellulose and its derivatives. J. Biomed. Mater. Res. 1989, 23, 125–133. [Google Scholar] [CrossRef]
  231. Ikada, Y. Biomedical applications of cellulose membranes. In Cellulose: Structural and Functional Aspects; Kennedy, J.F., Phillips, G.O., Williams, P.A., Eds.; Ellis Horwood: Chichester, UK, 1989; pp. 447–455. [Google Scholar]
  232. Barbié, C.; Chauveaux, D.; Barthe, X.; Baquey, C.; Poustis, J. Biological behaviour of cellulosic materials after bone implantation: Preliminary results. Clin. Mater. 1990, 5, 251–258. [Google Scholar] [CrossRef]
  233. Gross, U.; Muller-Mai, C.; Voigt, C. The tissue response on cellulose cylinders after implantation in the distal femur of rabbits. In Proceedings of the Fourth World Biomaterials Congress, Berlin, Germany, 19–24 May 1992; p. 192. [Google Scholar]
  234. Märtson, M.; Viljanto, J.; Hurme, T.; Saukko, P. Biocompatibility of cellulose sponge with bone. Eur. Surg. Res. 1998, 30, 426–432. [Google Scholar] [CrossRef]
  235. Bhatnagar, A.; Sain, M. Processing of cellulose nanofiber-reinforced composites. J. Reinf. Plast. Comp. 2005, 24, 1259–1268. [Google Scholar] [CrossRef]
  236. Fricain, J.C.; Granja, P.L.; Barbosa, M.A.; De Jéso, B.; Barthe, N.; Baquey, C. Cellulose phosphates as biomaterials. In vivo biocompatibility studies. Biomaterials 2002, 23, 971–980. [Google Scholar] [CrossRef]
  237. Svensson, A.; Nicklasson, E.; Harrah, T.; Panilaitis, B.; Kaplan, D.L.; Brittberg, M.; Gatenholm, P. Bacterial cellulose as a potential scaffold for tissue engineering of cartilage. Biomaterials 2005, 26, 419–431. [Google Scholar] [CrossRef] [PubMed]
  238. Märtson, M.; Viljanto, J.; Laippala, P.; Saukko, P. Connective tissue formation in subcutaneous cellulose sponge implants in the rat. Eur. Surg. Res. 1998, 30, 419–425. [Google Scholar] [CrossRef] [PubMed]
  239. Fundueanu, G.; Constantin, M.; Esposito, E.; Cortesi, R.; Nastruzzi, C.; Menegatti, E. Cellulose acetate butyrate microcapsules containing dextran ion-exchange resins as self-propelled drug release system. Biomaterials 2005, 26, 4337–4347. [Google Scholar] [CrossRef] [PubMed]
  240. Entcheva, E.; Bien, H.; Yin, L.; Chung, C.Y.; Farrell, M.; Kostov, Y. Functional cardiac cell constructs on cellulose-based scaffolding. Biomaterials 2004, 25, 5753–5762. [Google Scholar] [CrossRef]
  241. Tate, M.C.; Shear, D.A.; Hoffman, S.W.; Stein, D.G.; LaPlaca, M.C. Biocompatibility of methylcellulose-based constructs designed for intracerebral gelation following experimental traumatic brain injury. Biomaterials 2001, 22, 1113–1123. [Google Scholar] [CrossRef]
  242. Hoseini, S.M.; Khosravi-Darani, K.; Mozafari, M.R. Nutritional and medical applications of spirulina microalgae. Mini Rev. Med. Chem. 2013, 13, 1231–1237. [Google Scholar] [CrossRef]
  243. Granja, P.L.; Barbosa, M.A.; Pouységu, L.; De Jéso, B.; Rouais, F.; Baquey, C. Cellulose phosphates as biomaterials. Mineralization of chemically modified regenerated cellulose hydrogels. J. Mater. Sci. 2001, 36, 2163–2172. [Google Scholar] [CrossRef]
  244. Naseri-Nosar, M.; Salehi, M.; Hojjati-Emami, S. Cellulose acetate/poly lactic acid coaxial wet-electrospun scaffold containing citalopram-loaded gelatin nanocarriers for neural tissue engineering applications. Int. J. Biol. Macromol. 2017, 103, 701–708. [Google Scholar] [CrossRef]
  245. Wang, S.; Sun, C.; Guan, S.; Li, W.; Xu, J.; Ge, D.; Ma, X. Chitosan/gelatin porous scaffolds assembled with conductive poly (3, 4-ethylenedioxythiophene) nanoparticles for neural tissue engineering. J. Mat. Chem. B 2017, 5, 4774–4788. [Google Scholar] [CrossRef]
  246. Koide, S.S. Chitin-chitosan: Properties, benefits and risks. Nutr. Res. 1998, 18, 1091–1101. [Google Scholar] [CrossRef]
  247. Rinaudo, M. Chitin and chitosan: Properties and applications. Prog. Polym. Sci. 2006, 31, 603–632. [Google Scholar] [CrossRef]
  248. Madihally, S.V.; Matthew, H.W. Porous chitosan scaffolds for tissue engineering. Biomaterials 1999, 20, 1133–1142. [Google Scholar] [CrossRef]
  249. Ohkawa, K.; Cha, D.; Kim, H.; Nishida, A.; Yamamoto, H. Electrospinning of chitosan. Macromol. Rapid Commun. 2004, 25, 1600–1605. [Google Scholar] [CrossRef]
  250. Izzo, D.; Palazzo, B.; Scalera, F.; Gullotta, F.; Lapesa, V.; Scialla, S.; Gervaso, F. Chitosan scaffolds for cartilage regeneration: Influence of different ionic crosslinkers on biomaterial properties. Int. J. Polym. Mater. 2019, 68, 936–945. [Google Scholar] [CrossRef]
  251. Maged, A.; Abdelkhalek, A.A.; Mahmoud, A.A.; Salah, S.; Ammar, M.M.; Ghorab, M.M. Mesenchymal stem cells associated with chitosan scaffolds loaded with rosuvastatin to improve wound healing. Eur. J. Pharm. Sci. 2019, 127, 185–198. [Google Scholar] [CrossRef]
  252. Xue, Y.; Wu, M.; Liu, Z.; Song, J.; Luo, S.; Li, H.; Chen, F. In vitro and in vivo evaluation of chitosan scaffolds combined with simvastatin-loaded nanoparticles for guided bone regeneration. J. Mater. Sci. Mater. Med. 2019, 30, 47. [Google Scholar] [CrossRef]
  253. Skop, N.B.; Calderon, F.; Levison, S.W.; Gandhi, C.D.; Cho, C.H. Heparin crosslinked chitosan microspheres for the delivery of neural stem cells and growth factors for central nervous system repair. Acta Biomater. 2013, 9, 6834–6843. [Google Scholar] [CrossRef]
  254. Elnaggar, Y.S.R.; Etman, S.M.; Abdelmonsif, D.A. Intranasal Piperine-Loaded Chitosan Nanoparticles as Brain-Targeted Therapy in Alzheimer’s Disease: Optimization, Biological Efficacy, and Potential Toxicity. J. Pharm. Sci. 2015, 104, 3544–3556. [Google Scholar] [CrossRef]
  255. Raj, R.; Wairkar, S.; Sridhar, V.; Gaud, R. Pramipexole dihydrochloride loaded chitosan nanoparticles for nose to brain delivery: Development, characterization and in vivo anti-Parkinson activity. Int. J. Biol. Macromol. 2018, 109, 27–35. [Google Scholar] [CrossRef]
  256. Liu, S.H.; Ho, P.C. Intranasal administration of brain-targeted HP-β-CD/chitosan nanoparticles for delivery of scutellarin, a compound with protective effect in cerebral ischaemia. J. Pharm. Pharmacol. 2017, 69, 1495–1501. [Google Scholar] [CrossRef]
  257. Cooper, A.; Bhattarai, N.; Zhang, M. Fabrication and cellular compatibility of aligned chitosan–PCL fibers for nerve tissue regeneration. Carbohydr. Polym. 2011, 85, 149–156. [Google Scholar] [CrossRef]
  258. Kuo, Y.C.; Yeh, C.F.; Yang, J.T. Differentiation of bone marrow stromal cells in poly(lactide-co-glycolide)/chitosan scaffolds. Biomaterials 2009, 30, 6604–6613. [Google Scholar] [CrossRef] [PubMed]
  259. Marchesano, V.; Gennari, O.; Mecozzi, L.; Grilli, S.; Ferraro, P. Effects of lithium niobate polarization on cell adhesion and morphology. ACS Appl. Mater. Interfaces. 2015, 7, 18113–18119. [Google Scholar] [CrossRef] [PubMed]
  260. Tiffany, A.; Harley, B.A. Sequestration of biomolecules for controlled delivery in mineralized collagen scaffolds. In 42nd Society for Biomaterials Annual Meeting and Exposition 2019: The Pinnacle of Biomaterials Innovation and Excellence. Soc. Biomater. 2019, 40, 1526–7547. [Google Scholar]
  261. Sallent, I.; Capella-Monsonís, H.; Zeugolis, D.I. Production and Characterization of Chemically Cross-Linked Collagen Scaffolds. In Collagen; Humana Press: New York, NY, USA, 2019; pp. 23–38. [Google Scholar]
  262. Mohee, L.; Offeddu, G.S.; Husmann, A.; Oyen, M.L.; Cameron, R.E. Investigation of the intrinsic permeability of ice-templated collagen scaffolds as a function of their structural and mechanical properties. Acta Biomater. 2019, 83, 189–198. [Google Scholar] [CrossRef]
  263. Liu, Y.; Nelson, T.; Chakroff, J.; Cromeens, B.; Johnson, J.; Lannutti, J.; Besner, G.E. Comparison of polyglycolic acid, polycaprolactone, and collagen as scaffolds for the production of tissue engineered intestine. J. Biomed. Mater. Res. B 2019, 107, 750–760. [Google Scholar] [CrossRef]
  264. Lei, X.; Gao, J.; Xing, F.; Zhang, Y.; Ma, Y.; Zhang, G. Comparative evaluation of the physicochemical properties of nano-hydroxyapatite/collagen and natural bone ceramic/collagen scaffolds and their osteogenesis-promoting effect on MC3T3-E1 cells. Regen. Biomater. 2019, 6, 361–371. [Google Scholar] [CrossRef]
  265. Radhakrishnan, S.; Nagarajan, S.; Bechelany, M.; Kalkura, S.N. Collagen Based Biomaterials for Tissue Engineering Applications: A Review. In Processes and Phenomena on the Boundary between Biogenic and Abiogenic; Nature; Springer: Cham, Switzerland, 2020; pp. 3–22. [Google Scholar]
  266. Eltom, A.; Zhong, G.; Muhammad, A. Scaffold Techniques and Designs in Tissue Engineering Functions and Purposes: A Review. Adv. Mater. Sci. Eng. 2019. [Google Scholar] [CrossRef] [Green Version]
  267. Mohseni, M.; Castro, N.J.; Dang, H.P.; Nguyen, T.D.; Ho, H.M.; Tran, M.P.N.; Tran, P.A. Adipose tissue regeneration: Scaffold—Biomaterial strategies and translational perspectives. In Biomaterials in Translational Medicine; Academic Press: Cambridge, MA, USA, 2019; pp. 291–330. [Google Scholar]
  268. Ceballos, D.; Navarro, X.; Dubey, N.; Wendelschafer-Crabb, G.; Kennedy, W.R.; Tranquillo, R.T. Magnetically Aligned Collagen Gel Filling a Collagen Nerve Guide Improves Peripheral Nerve Regeneration. Exp. Neurol. 1999, 158, 290–300. [Google Scholar] [CrossRef]
  269. Dubey, N.; Letourneau, P.C.; Tranquillo, R.T. Guided Neurite Elongation and Schwann Cell Invasion into Magnetically Aligned Collagen in Simulated Peripheral Nerve Regeneration. Exp. Neurol. 1999, 158, 338–350. [Google Scholar] [CrossRef]
  270. Eguchi, Y.; Ogiue-Ikeda, M.; Ueno, S. Control of orientation of rat Schwann cells using an 8-T static magnetic field. Neurosci. Lett. 2003, 351, 130–132. [Google Scholar] [CrossRef]
Figure 1. Various mechanical stimuli exerted on the cell induce changes in plasma membrane tension, eliciting piezo-channel openings (adapted from [59]).
Figure 1. Various mechanical stimuli exerted on the cell induce changes in plasma membrane tension, eliciting piezo-channel openings (adapted from [59]).
Polymers 12 00161 g001
Figure 2. Classification of the nervous system.
Figure 2. Classification of the nervous system.
Polymers 12 00161 g002
Figure 3. Schematic illustration of the basic units of the nervous tissue: (A) neuron and (B) neuroglia.
Figure 3. Schematic illustration of the basic units of the nervous tissue: (A) neuron and (B) neuroglia.
Polymers 12 00161 g003
Figure 4. Potential difference in neural transmission as a function of time (adapted from [69]).
Figure 4. Potential difference in neural transmission as a function of time (adapted from [69]).
Polymers 12 00161 g004
Figure 5. Representative scheme of tissue regeneration in response to the mechanical and electrical stimulation on the piezoelectric scaffold.
Figure 5. Representative scheme of tissue regeneration in response to the mechanical and electrical stimulation on the piezoelectric scaffold.
Polymers 12 00161 g005
Figure 6. Scheme of permanent polarization in the α-helix.
Figure 6. Scheme of permanent polarization in the α-helix.
Polymers 12 00161 g006
Figure 7. Definition of the piezoelectric coefficients (adapted from [91]).
Figure 7. Definition of the piezoelectric coefficients (adapted from [91]).
Polymers 12 00161 g007
Figure 8. Structures of alpha and beta PVDF.
Figure 8. Structures of alpha and beta PVDF.
Polymers 12 00161 g008
Table 1. Natural polymers with the piezoelectric response.
Table 1. Natural polymers with the piezoelectric response.
Natural PolymersPiezoelectric Coefficient −d14 (pC/N)Ref.
CollagenSkin0.2[92]
Bone0.7[92]
Tendon2.0[93]
KeratinHorn1.8[94]
Wool0.1[94]
FibrinSalmon DNA0.07[93]
Table 2. Piezoelectric materials in nerve tissue engineering.
Table 2. Piezoelectric materials in nerve tissue engineering.
Material TypeScaffold DesignCells Type UsedRef.
Polyvinylidene Fluoride (PVDF)Film *Spinal cord neurons[104]
Film *Mouse neuroblastoma cells[105]
ChannelsMouse sciatic nerve model[106]
TubesWistar rats[107]
MembranesNeuronal cells[108]
FilmsStem cells[109]
Nanosheets *Rat neuronal cell line[110]
Fibers *Osteoblasts MG-63 cells[111]
FibersMesenchymal stem cells[112]
Poly[(vinylidene fluoride-co-trifluoroethylene]
(PVDF-TrFE)
FibersPoietics normal human neural progenitors[113]
Dorsal root ganglion[114]
FilmsPoietics normal human neural progenitors[113]
Membranes *Osteoblasts SaOS-2 cells[115]
TubesIn vivo implementation: rat sciatic nerves[116]
Fibers *Preosteoblasts[117]
Poly(3,4ethylenedioxythiophene)
(PEDOT)
FilmsFibroblast growth factor (bFGF)[118]
Films *-[119]
Films *Neural stem cells[120]
FilmsNeural stem cells[121]
Nanofibers *Brain neuroglioma cells[122]
Polylactic acid
(PLLA)
FibersSprague–Dawley rats[123]
PLLA blends for vascular differentiation in vitro[124]
Neural differentiation and growth in vitro[125,126]
PLLA blends for bone formation in vitro[127]
+PANi fibers *Nerve stem cells[128]
Poly(3-hydroxybutyrate-co-3-hydroxyvalerate)
(PHBV)
Fibers *Human mesenchymal stem cell[129]
CollagenFibersSchwann cells[130]
3D gel matricesEmbryonic rat cerebral cortices[131]
BaTiO3+PVDF matrixOsteoblasts[132]
* Tests conducted with electrical/mechanical stimulation.

Share and Cite

MDPI and ACS Style

Zaszczynska, A.; Sajkiewicz, P.; Gradys, A. Piezoelectric Scaffolds as Smart Materials for Neural Tissue Engineering. Polymers 2020, 12, 161. https://doi.org/10.3390/polym12010161

AMA Style

Zaszczynska A, Sajkiewicz P, Gradys A. Piezoelectric Scaffolds as Smart Materials for Neural Tissue Engineering. Polymers. 2020; 12(1):161. https://doi.org/10.3390/polym12010161

Chicago/Turabian Style

Zaszczynska, Angelika, Paweł Sajkiewicz, and Arkadiusz Gradys. 2020. "Piezoelectric Scaffolds as Smart Materials for Neural Tissue Engineering" Polymers 12, no. 1: 161. https://doi.org/10.3390/polym12010161

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop